butyl rubber-aliphatic polyester graft ... - citeseerx

159
Western University Scholarship@Western Electronic esis and Dissertation Repository June 2013 Butyl Rubber-Aliphatic Polyester Graſt Copolymers for Biomedical Applications: Synthesis and Analysis of Chemical, Physical and Biological Properties Bethany A. Turowec e University of Western Ontario Supervisor Dr. Elizabeth Gillies e University of Western Ontario Graduate Program in Biomedical Engineering A thesis submied in partial fulfillment of the requirements for the degree in Master of Engineering Science © Bethany A. Turowec 2013 Follow this and additional works at: hp://ir.lib.uwo.ca/etd Part of the Biomaterials Commons , Polymer Chemistry Commons , and the Polymer Science Commons is Dissertation/esis is brought to you for free and open access by Scholarship@Western. It has been accepted for inclusion in Electronic esis and Dissertation Repository by an authorized administrator of Scholarship@Western. For more information, please contact [email protected]. Recommended Citation Turowec, Bethany A., "Butyl Rubber-Aliphatic Polyester Graſt Copolymers for Biomedical Applications: Synthesis and Analysis of Chemical, Physical and Biological Properties" (2013). Electronic esis and Dissertation Repository. Paper 1313.

Upload: khangminh22

Post on 01-Apr-2023

1 views

Category:

Documents


0 download

TRANSCRIPT

Western UniversityScholarship@Western

Electronic Thesis and Dissertation Repository

June 2013

Butyl Rubber-Aliphatic Polyester GraftCopolymers for Biomedical Applications:Synthesis and Analysis of Chemical, Physical andBiological PropertiesBethany A. TurowecThe University of Western Ontario

SupervisorDr. Elizabeth GilliesThe University of Western Ontario

Graduate Program in Biomedical Engineering

A thesis submitted in partial fulfillment of the requirements for the degree in Master of Engineering Science

© Bethany A. Turowec 2013

Follow this and additional works at: http://ir.lib.uwo.ca/etd

Part of the Biomaterials Commons, Polymer Chemistry Commons, and the Polymer ScienceCommons

This Dissertation/Thesis is brought to you for free and open access by Scholarship@Western. It has been accepted for inclusion in Electronic Thesisand Dissertation Repository by an authorized administrator of Scholarship@Western. For more information, please contact [email protected].

Recommended CitationTurowec, Bethany A., "Butyl Rubber-Aliphatic Polyester Graft Copolymers for Biomedical Applications: Synthesis and Analysis ofChemical, Physical and Biological Properties" (2013). Electronic Thesis and Dissertation Repository. Paper 1313.

BUTYL RUBBER-ALIPHATIC POLYESTER GRAFT COPOLYMERS FOR BIOMEDICAL APPLICATIONS: SYNTHESIS AND ANALYSIS OF CHEMICAL,

PHYSICAL AND BIOLOGICAL PROPERTIES

(Thesis format: Monograph)

by

Bethany Turowec

Graduate Program in Biomedical Engineering

A thesis submitted in partial fulfillment of the requirements for the degree of

Master of Engineering Science

The School of Graduate and Postdoctoral Studies The University of Western Ontario

London, Ontario, Canada

© Bethany Turowec, 2013

ii

Abstract

Biomaterials can be used in a wide variety of medical applications owing to their breadth of

characteristics that can be imparted by varying their chemical structures. Butyl rubber (IIR),

which is a copolymer of isobutylene (IB) and small percentages of isoprene (IP), is

particularly attractive as a biomaterial because of its elastomeric mechanical properties,

biocompatibility, impermeability and high damping characteristics. IIR is typically

vulcanized through chemical-based crosslinking mechanisms. However, these methods are

not acceptable for biological applications. This thesis focuses on the synthesis of IIR-

polyester graft copolymers by grafting biodegradable and biocompatible polyesters including

poly(caprolactone) (PCL) and poly(D,L-lactide) (PDLLA) to the IIR backbone, and on the

study of their properties. These graft copolymers were synthesized by the grafting of amine-

terminated polyesters on a modified IIR backbone having activated carbonate moieties. The

resulting copolymers with varying polyester content were characterized by a wide range of

chemical techniques including nuclear magnetic resonance and infrared spectroscopic

methods as well as size exclusion chromatography. IIR-polyester copolymers displayed an

increase in Young’s modulus (E) and ultimate tensile strength (UTS) relative to IIR, while

maintaining cell-biomaterial interactions and non-toxicity. Despite significant polyester

content, the copolymers did not exhibit any significant degradation, even in 5 M NaOH at

37°C. Overall, this study reveals how the properties of IIR can be readily tuned through the

preparation of graft copolymers and provides a comprehensive evaluation of these properties

for their further study in biomedical applications.

Keywords

butyl rubber, biodegradable polymers, poly(caprolactone), poly(lactide), graft copolymer,

thermoplastic elastomer, ultimate tensile strength

iii

Acknowledgments

I would like to thank Dr. Elizabeth Gillies for her input, guidance and motivation throughout

my graduate studies. It was her patience, as well as continual optimism that drove me

through to completion of this project. Furthermore, the breadth of skills and knowledge that I

attained in the Gillies lab has increased my analytical thinking and problem-solving abilities

that will be applicable in future endeavours. I would also like to thank LANXESS Inc. for

their financial and research contributions. Although many LANXESS Inc. employees were

involved in this project, each individual’s input was greatly appreciated and provided a

positive input toward goal completion. Particular acknowledgment is given to Lorenzo

Ferrari for his continued support throughout the duration of the Gillies lab involvement with

LANXESS Inc.

Next, some of the greatest help and support received stemmed from fellow colleagues, both

past and present, of the Gillies lab. It was because of their knowledge and helpful attitudes

that not only provided the avenues necessary to complete this degree, but also made the

duration fun, positive and exciting. I would like to specifically thank Solmaz Karamdoust for

her constant help concerning any aspect of the butyl rubber project, as well as her and Rasoul

Soleimani for obtaining DSC data, Ryan Amos for cell culture training, Ryan McBride for

help analyzing and generating SEC traces for the degradation study and Aneta Borecki, the

lab technician, for collecting SEC data.

Thank you to the examiners, Dr. Amin Rizkalla, Dr. Joe Gilroy and Dr. JunYang that have

taken time to read and consider this thesis. I would also like to thank support staff at UWO:

Dr. Mat Willans (NMR), Dr. Heng-Yong Nie (AFM training), Dr. Paul Ragogna and group

(training and equipment usage), Nicole Bechard and Karen Nygard (Biotron microscopy),

chemstores staff and the Biomedical Engineering department.

Finally, I would like to thank my family and friends, but especially my parents, Sonia and

Tony Turowec. I am forever grateful for your resounding love, support and advice and

cannot imagine life without your blessings.

iv

Table of Contents

Abstract ............................................................................................................................... ii

Acknowledgments.............................................................................................................. iii

Table of Contents ............................................................................................................... iv

List of Tables .................................................................................................................... vii

List of Figures .................................................................................................................. viii

List of Schemes ................................................................................................................ xiv

List of Equations .............................................................................................................. xvi

List of Abbreviations ...................................................................................................... xvii

List of Appendices ........................................................................................................... xxi

1 Introduction .................................................................................................................... 1

1.1 Polymers and their Importance as Biomaterials ..................................................... 1

2 Background and Literature Review ............................................................................... 3

2.1 Polymers as Biomaterials ........................................................................................ 3

2.2 Natural Polymers .................................................................................................... 3

2.2.1 Collagen ...................................................................................................... 4

2.2.2 Chitosan ...................................................................................................... 5

2.2.3 Hyaluronic Acid .......................................................................................... 5

2.2.4 Applications ................................................................................................ 6

2.3 Synthetic Polymers ................................................................................................. 9

2.3.1 Biodegradable ............................................................................................. 9

2.3.2 PIB and IIR Copolymers ........................................................................... 17

2.4 Applications of IIR Materials ............................................................................... 28

2.5 Evaluation of Biomaterials ................................................................................... 31

2.5.1 Chemical Characterization ........................................................................ 31

v

2.5.2 Physical Characterization.......................................................................... 32

2.5.3 Biological Characterization ...................................................................... 45

2.6 Thesis Objectives .................................................................................................. 51

3 Results and Discussion ................................................................................................. 53

3.1 Synthesis and Chemical Characterization of IIR-PCL Copolymers ..................... 53

3.1.1 ROP of ε-caprolactone from the IIR Backbone ........................................ 53

3.1.2 Grafting of PCL onto the IIR Backbone ................................................... 57

3.2 Grafting of PDLLA onto the IIR Backbone ......................................................... 65

3.2.1 PDLLA Functionalization ......................................................................... 65

3.2.2 IIR and PDLLA Grafting .......................................................................... 68

3.3 Preparation of IIR-PCL/PDLLA Blends ............................................................... 71

3.4 Physical Characterization of Graft Copolymers ................................................... 72

3.4.1 Atomic Force Microscopy ........................................................................ 72

3.4.2 Water Contact Angle Measurements ........................................................ 78

3.4.3 Tensile Testing .......................................................................................... 80

3.5 Degradation Study of IIR-PCL/PLA Graft Copolymers ...................................... 88

3.5.1 Mass Evolution and Scanning Electron Microscopy ................................ 88

3.5.2 Size Exclusion Chromatography............................................................... 95

3.6 Bioassays and Compatibility................................................................................. 99

3.6.1 Cell Growth on Polymer Films ................................................................. 99

3.6.2 MTT Toxicity Assay ............................................................................... 103

4 Materials and Methods ............................................................................................... 105

4.1 General Procedures and Materials ...................................................................... 105

4.2 Graft Copolymer Synthesis and Chemical Characterization .............................. 106

4.2.1 Synthesis of Polymer 3.8a....................................................................... 106

4.2.2 Synthesis of Polymer 3.9a....................................................................... 107

vi

4.2.3 Synthesis of Polymer 3.12 ...................................................................... 108

4.2.4 Synthesis of Graft Copolymer 3.14 ........................................................ 108

4.3 Physical Characterization.................................................................................... 110

4.3.1 Atomic Force Microscopy ...................................................................... 110

4.3.2 Water Contact Angle............................................................................... 110

4.3.3 Scanning Electron Microscopy ............................................................... 110

4.3.4 Mechanical Testing of Graft Copolymers............................................... 110

4.3.5 Degradation Study .................................................................................. 111

4.4 Biological Characterization ................................................................................ 112

4.4.1 Cell Growth on Polymer Films ............................................................... 112

4.4.2 MTT Toxicity Assay ............................................................................... 113

5 Conclusions and Future Work .................................................................................... 115

References ....................................................................................................................... 117

Appendices ...................................................................................................................... 128

Curriculum Vitae ............................................................................................................ 136

vii

List of Tables

Table 2.1 – Air loss after automobile driving tests (Reproduced with permission from John

Wiley & Sons).109

.................................................................................................................... 21

Table 3.1 – Varying conditions to afford IIR and PCL graft copolymers by ROP of ε-

caprolactone from –OH moiety on IIR backbone. .................................................................. 54

Table 3.2 – PCL thermal properties (provided by Polymer SourceTM

). ................................. 58

Table 3.3 – IIR-PCL graft copolymers. .................................................................................. 63

Table 3.4 – PDLLA homopolymer and IIR-PDLLA graft copolymer: PDLLA content and

thermal properties. .................................................................................................................. 69

Table 3.5 – Contact angle of PCL and PDLLA homopolymers and copolymers................... 78

Table 3.6 – Tensile data of graft copolymers and IIR. ........................................................... 81

Table 3.7 – Tensile data of polymer blends. ........................................................................... 86

Table 4.1 – Modified protocol adapted from ISO 10993 Technical Committee. (2007).

International Organization for Standardization (ISO) .......................................................... 114

viii

List of Figures

Figure 2.1 – Common tripeptide sequence of collagen composed of glycine (Gly), proline

(Pro) and hydroxyproline (Hyp) leading to helical structure. (Reprinted from Progress in

Polymer Science, 35/4, Puppi et. al., Polymeric materials for bone and cartilage repair (403-

440). Copyright (2010), with permission from Elsevier. .......................................................... 4

Figure 2.2 – Structure of chitosan. ............................................................................................ 5

Figure 2.3 – Chemical structure of hyaluronic acid. ................................................................. 6

Figure 2.4 – Structural representations of diisocyanates. ....................................................... 10

Figure 2.5 – Chemical structures of various POEs. ................................................................ 11

Figure 2.6 – Chemical structures of PDLA (S-enantiomer), PLLA (R-enantiomer), PDLLA

(racemic mixture) and PCL. .................................................................................................... 14

Figure 2.7 – Molecular weight changes for a porous PCL structure (triangle) and a linear

PCL structure (square). Reproduced with permission from Taylor & Francis, 2007. ............ 16

Figure 2.8 – Commercialized IIR production: general IIR slurry polymerization (Reproduced

with permission from John Wiley & Sons, 1990).109

.............................................................. 19

Figure 2.9 – Isoprene unit enters chain predominantly in trans-1,4 configuration. ................ 19

Figure 2.10 – Monomer-concentration dependence on number of active sites and styrene

conversion (left); monomer-concentration dependence on MW of grafted PS and MW of PS

homopolymer (Biomaterials, 29, 2008, 448-460 © 2007 Elsevier Ltd. with kind permission

from Springer Science and Business Media). ......................................................................... 27

Figure 2.11 – Fluorescence confocal microscopy images following adsorption of a

rhodamine-fibrinogen conjugate. PEO content: a) 2%, b) 4%, c) 6%, d) 12%, e) 24% and f)

34% (Reprinted with permission from Macromolecules 2011, 44, 6405. Copyright (2011)

American Chemical Society). ................................................................................................. 29

ix

Figure 2.12 – Cartoon representation of PIB-b-PS showing elastomeric entanglements (PIB)

and hard segments (PS) (Reprinted from Biomaterials, 29/4, Pinchuk et. al., Medical

applications of poly(styrene-block-isobutylene-block-styrene) (“SIBS”) (448-460). Copyright

(2008), with permission from Elsevier). ................................................................................. 30

Figure 2.13 – Contact angle measurement (θc) and interphase-energy between 3 phases

(values in Young’s equation found below). ............................................................................ 33

Figure 2.14 – Temperature dependence of stiffness of typical thermoplastic elastomers

(Reprinted from Handbook of Thermoplastic Elastomers, 1st Edition, Drobny, Jiri George,

Introduction (1-7). Copyright (2007), with permission from Elsevier).58

.............................. 37

Figure 2.15 – Instron (3300 series) tensile testing instrument. ............................................... 39

Figure 2.16 – Definitions of uniaxial stress, strain and elastic deformation. (Reprinted from

Nanomaterials, Nanotechnologies and Design: An Introduction for Engineers and Architects,

1st Edition, Ashby, Michael F., Chapter 4 – Material Classes, Structure and properties.

Copyright (2009), with permission from Elsevier).189

............................................................ 41

Figure 2.17 – Stress strain curve for a polymer. Definitions of uniaxial stress, strain and

elastic deformation. (Reprinted from Nanomaterials, Nanotechnologies and Design: An

Introduction for Engineers and Architects, 1st Edition, Ashby, Michael F., Chapter 4 –

Material Classes, Structure and properties. Copyright (2009), with permission from

Elsevier).189

............................................................................................................................. 42

Figure 2.18 – Material property charts showing: a) Young’s modulus and its relation to

material density; b) Young’s modulus and material strength to defined the yield strain (σy/E),

where a material no longer behaves elastically. (Reprinted from Nanomaterials,

Nanotechnologies and Design: An Introduction for Engineers and Architects, 1st Edition,

Ashby, Michael F., Chapter 4 – Material Classes, Structure and properties. Copyright (2009),

with permission from Elsevier).189

.......................................................................................... 44

Figure 2.19 – PCL and PCL-Col (collagen) materials: confocal scanning laser microscopies

showing differences in cell proliferation on different surfaces (Reprinted from Biomaterials,

x

25/11, Cheng and Teoh, Surface modification of ultra thin poly (ε-caprolactone) films using

acrylic acid and collagen (1991-2001). Copyright (2003), with permission from Elsevier). . 46

Figure 2.20 – Morphologies of hMSCs cultured on various substrates: B) glass control; C)

PLCL; D) AAc-PLCL and E) gelatin-AAc-PLCL. Scale bar = 200 μm (Reprinted with

permission from Shin, Y. M.; Kim, K.-S.; Lim, Y. M.; Nho, Y. C.; Shin, H.

Biomacromolecules 2008, 9, 1772. Copyright 2008 American Chemical Society). .............. 47

Figure 2.21 – (a-f) Myoblasts stained for vinculin (light speckles), F-actin (outer periphery)

and nuclei, scale bar = 50 μm. (a) 4 h after seeding on glass coverslip control, (b) 4 h after

seeding on C500 surface, (c) 4 h after seeding on C6000 surface, (d) 12 h after seeding on

glass coverslip control, (e) 12 h after seeding on C500 surface, and (f) 12 h after seeding on

C6000 surface (Reprinted with permission from Dugan, J. M.; Gough, J. E.; Eichhorn, S. J.

Biomacromolecules 2010, 11, 2498, Copyright 2010 American Chemical Society). ............ 48

Figure 2.22 – Confocal microscopy images of C2C12 cells adhered to control and copolymer

surfaces: a) glass (control); b) IIR (control); c) 18wt% PEO ; d) 32wt% PEO; e) 65 wt%

PEO; f) 83 wt% PEO. Nuclei (dark inner portion) and F-actin fibres (lighter periphery) with

image area = 0.22 x 0.22 mm (Reprinted with permission from John Wiley & Sons, 2013). 49

Figure 3.1 – NMR spectra showing: a) hydroxylated IIR; b) 50 wt% IIR-PCL graft

copolymer following methanol precipitation (label k denotes terminal methylene PCL); c)

second precipitation in acetone of the same polymer from b), confirming a decrease in PCL

content to 16 wt% (label k denotes terminal PCL methylene). .............................................. 55

Figure 3.2 – SEC trace for 50 wt% PCL shows free homopolymer. 50 wt% PCL (post-

purification) trace reveals homopolymer removal achieved with secondary precipitation.

Detection was based on differential refractive index. ............................................................. 56

Figure 3.3 – PCL (900 g/mol) functionalization (with k referring to the terminal methylene):

a) 3.7a; b) 3.8a; c) 3.9a. .......................................................................................................... 59

Figure 3.4 – SEC traces of PCL derivatives throughout the functionalization process: a) 3.7a

– 3.9a (900 g/mol); b) 3.7b – 3.9b (3500 g/mol). ................................................................... 59

xi

Figure 3.5 – 1H NMR spectra (CDCl3, 600MHz) of a) activated IIR; b) copolymer 3.15; and

c) copolymer 3.16 showing how PCL content can be determined based on the relative

intensities of PCL to PIB, as well as reaction conversion determination based on peaks from

4.8-5.3 ppm. ............................................................................................................................ 62

Figure 3.6 – SEC traces for: ungrafted IIR (3.1) and each IIR-PCL graft copolymer (3.14-

3.16). ....................................................................................................................................... 63

Figure 3.7 – Schematic depicting PDLLA functionalization: a) 3.10 (2800 g/mol), starting

material; b) 3.11, t-BOC protected β-alanine derivative; c) 3.12, amine-liberated derivative.

In 3.10-3.12, f represents the terminal methylene. ................................................................. 66

Figure 3.8 – SEC traces elucidating functionalization of PDLLA 3.10-3.12. ........................ 67

Figure 3.9 – PDLLA content in copolymer 3.17: determined via integration corresponding to

PDLLA multiplet from 5.13-5.23 ppm and PIB singlet at 1.41 ppm. .................................... 69

Figure 3.10 – SEC traces of ungrafted IIR (polymer 3.1) and IIR-PDLLA graft copolymer

(3.17). ...................................................................................................................................... 70

Figure 3.11 – Topography of copolymers: a) 3.14; b) 3.15; c) 3.16; d) 3.17. ........................ 73

Figure 3.12 – Phase contrast of copolymers: a) 3.14; b) 3.15; c) 3.16; d) 3.17. ..................... 74

Figure 3.13 – AFM analysis of copolymer 3.16. Before annealing: a) topography; b) phase

contrast. After annealing: c) topography; d) phase contrast. .................................................. 75

Figure 3.14 – Topography images of IIR-polyester blends: a) 15 wt% PCL; b) 32 wt% PCL;

c) 44 wt% PCL; d) 30 wt% PDLLA. ...................................................................................... 76

Figure 3.15 – Phase contrast of polymer blends: a) 15 wt% PCL; b) 32 wt% PCL; c) 44 wt%

PCL; d) 30 wt% PDLLA......................................................................................................... 77

Figure 3.16 – Stress-strain curve for: a) IIR; b) copolymer 3.14; c) copolymer 3.15; d)

copolymer 3.16; e) copolymer 3.17. ....................................................................................... 81

xii

Figure 3.17 – Stress vs Strain of a variety of materials (with permission to reprint from MIT

OpenCourseWare, http://flic.kr/p/66XeQc). ........................................................................... 82

Figure 3.18 – Thermoplastic PURs with varying soft segment contents. (Reprinted from

Polymer, 49/19, Xu et. al., Morphology and properties of thermoplastic polyurethanes with

dangling chains in ricinoleate-based soft segments (4248-4258). Copyright (2008), with

permission from Elsevier).244

.................................................................................................. 83

Figure 3.19 – Toughness-Modulus plot for current implant materials and the mechanical

regions for orthopedic hard tissue, orthopedic soft tissue, and cardiovascular tissue.

(Reprinted with permission from Taylor & Francis, 2013).246

............................................... 85

Figure 3.20 – Stress-strain curve for: a) 15 wt% PCL blend; b) 32 wt% PCL blend; c) 44

wt% PCL blend; d) 30 wt% PDLLA blend. ........................................................................... 85

Figure 3.21 – Mass loss of copolymers 3.14-3.16, PCL and IIR controls. ............................. 89

Figure 3.22 – Mass loss of copolymer 3.17 as well as PDLLA and IIR controls................... 90

Figure 3.23 – SEM imaging taken at 100X magnification under variable pressure mode at To

and 4 months of: a) and b) IIR; c) and d) copolymer 3.14; e) and f) copolymer 3.15; g) and h)

copolymer 3.16. ...................................................................................................................... 91

Figure 3.24 – SEM imaging taken at 100X magnification under variable pressure mode at To

and 2 months of: a) and b) copolymer 3.17. Macroscopic images of To-2 months of: c)

copolymer 3.17. ...................................................................................................................... 92

Figure 3.25 – Successive macroscopic images representing To, 1 month, 2 months, 3 months

and 4 months of: a) IIR control; b) copolymer 3.14; c) copolymer 3.15; d) copolymer 3.16. 93

Figure 3.26 – MW data for IIR control as well as copolymer 3.14, 3.15, 3.16 and 3.17: a)

change in Mn; b) change in Mw. .............................................................................................. 95

Figure 3.27 – MW profiles elucidating each time point over the 4 month study period for: a)

IIR control; b) copolymer 3.14; c) copolymer 3.15; d) copolymer 3.16. ............................... 96

xiii

Figure 3.28 – MW profile elucidating each time point over the 2 month study period for

copolymer 3.17. ...................................................................................................................... 97

Figure 3.29 – Growth of murine myoblast cells on: a) glass; b) IIR; c) PCL; d) PDLLA Cell

imaging of copolymers: e) 3.14; f) 3.15; g) 3.16; h) 3.17. .................................................... 100

Figure 3.30 – Cell adhesivity quantified by determining average cell concentrations for each

substrate. ............................................................................................................................... 102

Figure 3.31 – MTT cytotoxicity assay performed on: a) graft copolymers 3.14-3.16; b) graft

copolymer 3.17. .................................................................................................................... 104

xiv

List of Schemes

Scheme 2.1 – Hydrolysis of PCL to 6-hydroxylcaproic acid and acetyl coenzyme A

intermediates followed by elimination from the body through the citric acid cycle. ............. 12

Scheme 2.2 – Mechanism of stannous octoate polymerization of PCL: 1/2 – formation of

stannous alkoxide initiator; 3 – deactivation of catalyst; 4 – coordination/insertion of

monomer; 5 – chain transfer of active polymerizing centre to alcohol. ................................. 13

Scheme 2.3 – Cationic polymerization of IB governed by initiation, propagation and

termination. ............................................................................................................................. 20

Scheme 2.4 – Vulcanization of IIR: a) sulfur-based crosslinking; b) dioxime curing; c)

general structure of resin capable of vulcanization. ................................................................ 23

Scheme 2.5 – IIR bromination followed by isomerization and HX elimination. ................... 24

Scheme 2.6 – Reaction schematic elucidating SIBS production via bifunctional HDCE

initiatior. .................................................................................................................................. 25

Scheme 2.7 – Synthesis of IIR-PEO graft copolymers. .......................................................... 26

Scheme 3.1 – Epoxidation followed by hydroxylation of IIR with 2.2mol% IP and

subsequent ROP of ε-caprolactone from IIR backbone. ......................................................... 53

Scheme 3.2 – Synthesis of t-BOC-protected β-alanine anhydride. ........................................ 57

Scheme 3.3 – Functionalization of PCL (3.7a/b) by first reacting with BOC-protected β-

alanine (3.6) to produce the protected derivative (3.8a/b), followed by deprotection with TFA

(n = 8 for 900 g/mol PCL, 3.7a, initiated with ethylene glycol derivative and n = 31 for 3500

g/mol PCL, 3.7b initiated with ethanol).................................................................................. 58

Scheme 3.4 – p-nitrophenyl chloroformate (PNPC) activated rubber synthesis. ................... 61

Scheme 3.5 – Synthesis of PCL graft copolymers: 3.14 – 15 wt% PCL (n=8); 3.15 – 32 wt%

PCL (n=31); 3.16 – 44 wt% PCL (n=31). ............................................................................... 61

xv

Scheme 3.6 – Functionalization of 3.10 by first reacting with BOC-protected β-alanine (3.6)

to afford 3.11, followed by deprotection with TFA (3.12). .................................................... 65

Scheme 3.7 – Synthesis of PDLLA Graft copolymer: 3.17 – 30 wt% PDLLA. .................... 68

xvi

List of Equations

Equation 1 – Definition of polydispersity index where Mn is the total weight of the sample

divided by the number of molecules (arithmetic mean) and Mw fairly accounts for the

contributions of different sized chains. ................................................................................... 32

Equation 2 – where ϴ = measured contact angle and γ is the surface tension of the solid-gas

(SG), solid-liquid (SG) and liquid-gas (LG) interface. ........................................................... 33

Equation 3 – Young’s modulus determination for a material at a given strain and stress. ..... 39

Equation 4 – Linear relationship relating the shear strain, γ to the shear stress, τ. ................. 40

Equation 5 – Linear relationship showing proportionality between the dilatation, Δ and

pressure, p. .............................................................................................................................. 40

Equation 6 – Definition of Poisson’s ratio where εt is the transverse strain and ε is the axial

strain. ....................................................................................................................................... 40

Equation 7 – Relation between Young’s modulus (E), shear modulus (G) and bulk modulus

(K). .......................................................................................................................................... 41

Equation 8 – Definition of plastic strain. ................................................................................ 42

Equation 9 – Where = average mass of initial 3 discs for a given time point and =

average mass of final 3 discs for a given time point. ............................................................ 111

xvii

List of Abbreviations

(1H) NMR hydrogen nuclear magnetic resonance spectroscopy

AFM atomic force microscopy

BDR butadiene rubber

BIIR bromo-butyl rubber

BOC di-tert-butyl dicarbonate

br. s broad singlet

CA cyanoacrylate

CIIR chloro-butyl rubber

DAPI 4’-6-diamidino-2-phenylindole dyhydrocholoride

DCM dichloromethane

DES drug eluting stent

DMEM Dulbecco’s Modified Eagle Medium

DPNR deproteinized natural rubber

DSC differential scanning calorimetry

E Young’s modulus

ECM extracellular matrix

FTIR fourier transform infrared spectroscopy

G shear modulus

GPC gel permeation chromatography

HA hyaluronic acid

HSD honestly significant difference

IB isobutylene

IIR butyl rubber

IP isoprene

xviii

J coupling constant

K bulk modulus

m-CPBA meta-chloroperoxybenzoic acid

Mn number-average molecular weight

MSA methanesulfonic acid

MTT 3-(4,5-Dimethylthiazol-2-yl)-2,5-Diphenyltetrazolium Bromide

Mw weight-average molecular weight

MW molecular weight

Mz z-average molecular weight

Mν viscosity molecular weight

NBR acrylonitrile butadiene rubber

NR natural rubber

p pressure

PBD poly(butadiene)

PBS phosphate buffered saline

PCL poly(caprolactone)

PDI polydispersity index

PDLA poly(D-lactide)

PDLLA poly(D,L-lactide)

PEG poly(ethylene glycol)

PEO poly(ethylene oxide)

PIB poly(isobutylene)

PLLA poly(L-lactide)

PMMA poly(methyl methacrylate)

PNPC 4 (para)-nitrophenyl chloroformate

xix

POE poly(ortho esters)

PS poly(styrene)

PUR poly(urethane)

PVC poly(vinyl chloride)

RB-402 butyl rubber (LANXESS Inc. derivative)

ROP ring-opening polymerization

RT room temperature

SBR styrene butadiene rubber

SDS sodium lauryl sulfate

SEC size exclusion chromatography

SEM scanning electron microscopy

SIBS styrene-isobutylene-styrene

SMP shape memory polymer

TE tissue engineering

TFA trifluoroacetic acid

Tg glass transition temperature

THF tetrahydrofuran

Tm melt transition temperature

TPE thermoplastic elastomer

UTS ultimate tensile strength

WCA water contact angle

wt% weight percentage

γ shear strain

Δ dilatation

ε strain

xx

εb elongation at break

εt lateral strain

ν Poisson’s ratio

σ stress

σel elastic limit

σy yield strength

τ shear stress

xxi

List of Appendices

Appendix A – 1H NMR depicting functionalization of 3.7b: a) initial homopolymer; b)

reacted with t-BOC-protected β-alanine (3.8b); c) and deprotected with TFA (3.9b). ........ 128

Appendix B – 1H NMR showing PCL content of: a) copolymer 3.16 (44 wt% PCL) and b)

copolymer 3.15 (32wt% PCL). ............................................................................................. 129

Appendix C – FTIR of: a) 3.8a and 3.9a; b) 3.11 and 3.12. ................................................. 130

Appendix D – FTIR of: a) copolymer 3.14; b) copolymer 3.15. .......................................... 131

Appendix E – FTIR of: a) copolymer 3.16; b) copolymer 3.17. .......................................... 132

Appendix F – DSC traces of: IIR-PCL copolymer after 1st precipitation; IIR-PCL copolymer

after a 2nd

precipitation into acetone, ridding of free homopolymer as evidenced by

disappearance of the second Tm at 53°C. ............................................................................. 133

Appendix G – DSC trace depicting the two Tgs of copolymer 3.17. .................................... 133

Appendix H – DSC traces depicting Tg and Tm of IIR-PCL graft copolymers: a) copolymer

3.14; b) copolymer 3.15; c) copolymer 3.16. ........................................................................ 134

Appendix I – DSC traces depicting Tg and Tm of IIR-PCL/PLA blends. ............................. 134

Appendix J – Raw data coinciding with Figure 3.21 for the degradation study of IIR-PCL

copolymers and IIR control. ................................................................................................. 135

Appendix K – Raw data coinciding with Figure 3.22 for the degradation study of IIR-

PDLLA copolymer and IIR control. ..................................................................................... 135

1

Chapter 1

1 Introduction

1.1 Polymers and their Importance as Biomaterials

Biomaterials possess the ability to perform with an appropriate host response in specific

applications.1 They are used extensively in a wide variety of applications, ranging from

cardiovascular, dental and neural implants to orthopaedic prosthetics and drug delivery

systems. Biomaterials have always been important as vehicles for the treatment of

disease, thereby constantly improving health care. Early examples dating back thousands

of years ago are metals and wood for teeth replacement and glass for eyeball prosthetics.

However, the discovery of synthetic polymers such as poly(methacrylates) and

poly(urethanes) led to a much broader range of application possibilities. Moreover,

naturally occurring materials such as collagen are also being developed for better

applicability in biological systems.

Although biomaterials perhaps play a role in the lives of many individuals, there

are several difficulties involved with their use. The main issues stem from deficiencies in

understanding physical, chemical and biological responses that a given biomaterial may

elicit in biological systems, as well as the lack of proper performance in a specific

application.2 Since many biomaterials were not originally designed to be used in a

clinical setting, taking off-the-shelf products has proved to be problematic.3 Examples of

such consist of dialysis tubing derived from cellulose acetate, Dacron for synthesis of

vascular grafts and poly(urethane) for the fabrication of artificial hearts. The

aforementioned applications were unsuccessful as cellulose acetate caused platelet

activation, Dacron grafts were limited to large-diameter vessel applications and

poly(urethane) did not supply sufficient blood-material interactions, respectively.4

In order to reduce issues related to using materials in applications for which they

were not specifically designed, research has been directed toward modifying chemical

structures to improve their mechanical properties, degradability and biocompatibility.5

Modern biomaterial production involves thorough understanding of cell-polymer

2

interactions in order to prevent or minimize undesirable cellular responses.6 Applications

such as polymer-coated stents for drug, protein and hormone delivery, tissue engineering

(TE) and other polymer/cell combinations such as artificial corneas, cartilage and bone,

makes it very important to understand biological response.5 Biomaterials have made great

impacts on medicine and current technology and will therefore impact biomedical

application advancements. As the aging population of developed countries continues to

grow, the demand for biomedical products to enhance life quality and longevity will

proportionally increase.

The focus of this thesis will be on preparing new potential biomaterials based on butyl

rubber (IIR)-polyester copolymers. The incorporation of an elastomeric component in the

biomaterial is especially important when considering its ability to mimic soft tissues.

Because humans predominantly consist of soft tissues, biomaterials that possess similar

mechanical and viscoelastic properties have potential for application in a multitude of

areas ranging from vascular prostheses (blood-interfacing implants) to breast implants

(non-blood-interfacing). Although elastomers are attractive due to their compliance with

soft or cardiovascular tissues, mechanical property enhancements may be required for

various applications. Therefore, in order to provide strength and rigidity, the

incorporation of polyesters (hard phase) with IIR (soft phase) will provide physical

crosslinks. The properties presented by these copolymers may be analogous to

thermoplastic elastomers (TPE). TPEs are typically composed of a phase which is hard at

ambient temperature, while the other is elastomeric. Phases are most commonly bonded

chemically through block/graft copolymerization.7 Without the hard phase, the elastomer

phase would flow freely, rendering it unsuitable for biomedical applications requiring

rigidity. For example, elastin and collagen are important components of various arteries.

Although functions of such soft tissues vary, it is the combination of the elastomeric

elastin with harder collagen that provides appropriate mechanical properties. With this

understanding, it will be interesting to investigate how the properties of IIR can be varied

chemically, physically and biologically to afford potential biomaterials.

3

Chapter 2

2 Background and Literature Review

2.1 Polymers as Biomaterials

Polymers are particularly interesting for use as biomaterials; their chemical, physical and

biological properties vary across a wide variety of structures, rendering them useful in

many different applications. In addition, through modification of the polymer backbone,

these properties can be specifically tuned. However, it is important to understand how

each polymer or polymeric system will impact its performance. Since applications can

range from soft tissue/organ replacement, drug delivery, wound dressings, to even

reconstruction of bone deficiencies, investigating a range of varying polymers to facilitate

new materials for biomedical applications has become increasingly apparent.

2.2 Natural Polymers

Biomedical applications involving the use of natural polymers such as collagen, chitosan

and alginate date back thousands of years. Natural polymers possess the obvious

advantages associated with biocompatibility; they do not elicit inflammatory responses or

other unsuitable side-effects that may result through the use of synthetic systems.

Although synthetic polymers are desirable as their properties can be tuned and controlled

with ease, the inherent issues stemming from biocompatibility still present natural

polymers as viable candidates for use as biomaterials.8 There are a wide variety of natural

polymers, leading to different avenues and applications. Proteins, such as collagen and

silk fibroin; polysaccharides, such as chitosan, hylauronic acid, alginates, dextrans and

starch-based materials; and microbial polyesters, such as polyhydroxyalkanoates are all

viable natural materials that are under current investigation for biomedical usage.

However, focus will be directed toward some notable materials including collagen,

chitosan and hyaluronic acid.

4

2.2.1 Collagen

Collagen is the most abundant protein in the body, which is owed to its high composition

in both skin and other musculoskeletal tissues.9 Type-I collagen (skin, tendon and bone)

is the most prevalent in mammals, providing the structural integrity and architecture.8

Type-I is composed of three polypeptide subunits consisting of similar amino acid

compositions: 33% glycine (Gly), 25% proline (Pro) and 25% hydroxyproline (Hyp).

These subunit chains allow collagen to undergo transcription, translation and post-

translational modification processes in fibroblasts and osteoblasts. Since these amino acid

subunits form polypeptides in typical sequences, it causes collagen to have a helical

structure, thereby providing it with its mechanical strength and resiliency (Figure 2.1).10

Moreover, its flexibility can be tuned by increasing the glycine content if it is required for

a specific application. It is because of this mechanical strength that utilizing collagen for

biomedical applications such as scaffolds,11,12

drug-delivery systems,13,14

shields for

contact lenses,15

sponges,16

hydrogels,17,18

nanoparticles19,20

and skin replacements,21,22

is

advantageous. In addition, its low antigenicity and good cell-binding properties make it

attractive for TE applications.23,24

Collagen sponges have been fabricated for cell and

tissue attachment,25,26

and to also enhance bone formation due to osteoblast

differentiation.27,28

However, because it requires crosslinking agents for certain

applications, this may render it unsuitable due to toxic byproducts.

Figure 2.1 – Common tripeptide sequence of collagen composed of glycine (Gly),

proline (Pro) and hydroxyproline (Hyp) leading to helical structure. (Reprinted

from Progress in Polymer Science, 35/4, Puppi et. al., Polymeric materials for bone

and cartilage repair (403-440). Copyright (2010), with permission from Elsevier.

5

2.2.2 Chitosan

Chitosan is a linear polyelectrolyte copolymer, which is composed of randomly

distributed 2-acetamido-2-deoxy-β-D-glucopyranose and N-acetyl-D-glucosamine

(chitin) units. Most units in the copolymer consist of the deacetylated version (2-amino-

2-deoxy-β-D-glucopyranose) making it hydrophilic thereby promoting cell adhesion,

proliferation and differentiation (Figure 2.2). Chitosan, like most natural polymers, is

biocompatible, and possesses other desirable properties including high charge density,

non-toxicity and mucoadhesion, rendering it appropriate for pharmaceutical and cosmetic

applications.8 The chain of chitosan is somewhat stiff, stabilizing a liquid crystalline

phase in acetic acid.29

The predominantly explored applications of chitosan involve non-

viral gene delivery due to its cationic nature allowing complex formation with DNA

molecules.30-32

Moreover, chitosan-based products are also appropriate for the delivery of

chemotherapeutics such as antibiotics, antiparasitics, anaesthetics and painkillers, via

routes involving injectable chitosan hydrogels. Chitosan’s novel properties make it an

appropriate natural polymer for property modification to result in a viable biomaterial for

cell therapy, TE and gene therapy. These TE applications include skin, bone, cartilage,

liver, nerve and blood vessel.33,34

However, its chemical modification cannot correct

deficiencies concerning mechanical weakness and instability, incapacity to maintain a

predefined shape, as well as impurities affecting material properties.35

Figure 2.2 – Structure of chitosan.

2.2.3 Hyaluronic Acid

Hyaluronic acid (HA) is an example of a polysaccharide, which consists of a high

molecular weight (MW) and linear backbone. Typically referred to as hyaluronan, this

polymer exists as a polyanion with alternating disaccharide units of β-1,3-N-acetyl-D-

6

glucosamine and glucuronic acid (Figure 2.3). HA is the main component of the

extracellular matrix (ECM). Not only does it serve as structural support, but it also

interacts with proteins, proteoglycans and other bioactive molecules, thereby contributing

to regulating processes such as cell behaviour, inflammation, angiogenesis and healing.36

Again, good biocompatibility as well as viscoelastic properties render HA attractive for

delivery systems, cell encapsulation; but most notably for TE due to its availability and

chain size manipulation. Although suitable for ECM remodeling because of cellular

interactions, its hydrophilicity does not favour cell attachment and tissue formation. In

order to alleviate these deficiencies, conjugation to collagen and fibronectin can be

performed to improve cellular interactions.37

Figure 2.3 – Chemical structure of hyaluronic acid.

2.2.4 Applications

Natural polymers are somewhat limited by the properties that they possess. In order to

modify their properties, they must be changed in a way that does not incorporate

synthetic materials. The advantage of natural polymers is their ability to interact within

biological systems in a reproducible and predictive manner; changing their properties

through a synthetic means may render these advantages obsolete. Chitosan has been

found interesting for a variety of applications, primarily owed to its degradation and

solubility, which can be tuned by substituting isobutyl at deacetylated sites without

altering its bioactivity.38

Chitosan does not exhibit foreign body reactions, thus

minimizing inflammatory responses, making it attractive for a range of in vivo

applications.39

For stent applications, there currently is one major contribution that

consists of a self-expanding chitosan stent.40

The stent employs a highly de-acetylated

version (slower degradation),41

implanted into the vas deferens of rats, displaying

7

adequate self-expansion. Finally, chitosan has the ability to form a high charge density in

weakly acidic solutions, producing cationic polymers. As a result, it can interact with

anionic polymers, negatively charged mucous membranes and DNA making it applicable

for mucoadhesives, bioadhesive drug delivery systems and for non-viral gene delivery

vehicles, respectively.42-44

Although collagen does not interact with anionic materials, its properties,

including enzymatic degradability and unique physico-chemical, mechanical and

biological properties, make it an interesting material for a variety of biomedical

applications.9 Collagen is the main component of the extracellular matrix, presenting it as

a strong candidate for TE/engineering applications. Because it is natural and abundant

amongst biological systems, it acts as a substrate for cell attachment, proliferation and

differentiation. Many applications involving spongy collagen matrices (Promogran®

),45

wound dressing materials (Biobrane® and Alloderm

®)46

and bilayer skin substitutes

(Integra® Dermal Regeneration Template)

47 have been developed and FDA approved for

treatment of ulcer wounds and thermal injuries. In addition, collagen has also been

investigated for usage as delivery vehicles for small molecule drugs. Current products

(Sulmycin®

-Implant, Collatamp®-G and Septocoll

®) focus on delivery of the antibiotic

gentamicin, resulting in prolonged local exposure with minimal systemic infiltration.48

However, the main source of collagen for biomedical applications originates from

bovine, porcine and equine skin or Achilles tendons. The wide-spread use of these

collagen-based biomaterials is therefore unforeseeable because of deficiencies and

variations associated with immune responses to materials derived from different species.

In order to make these applications realizable, human-sequenced collagen will have to be

recombinantly developed.49

Lastly, HA has presented itself as a unique biomaterial due to its structure; it is a

polysaccharide that is found in most, if not all, vertebrate tissues. HA is versatile, which

can be owed to its varying roles in biological processes, including cell migration and

differentiation control during embryogenesis, extracellular matrix organization and

metabolism and regulation for wound healing and inflammation.50

Because of these

properties, HA has predominantly been studied for wound dressing, TE and drug delivery

8

applications. More specifically, HA promotes angiogenesis, thereby reducing

inflammation51

and its high degree of chemical functionality allows it to undergo

crosslinking,52

making it appropriate for regulating wound sites and to tailor its

degradation rate for drug delivery. However, physical and biological limitations have

made HA impractical as a biomaterial due to its low water solubility, rapid resorption,

short residence time and anionic surface thereby prohibiting cellular attachment and

tissue formation.53

9

2.3 Synthetic Polymers

2.3.1 Biodegradable

Biodegradable synthetic polymers have been receiving considerable interest as

biomaterials; long-term biocompatibility issues with permanent implants has turned the

focus towards temporary therapeutic devices for a variety of applications.9 TE scaffolds,

regenerative medicine, drug eluting stents for controlled drug delivery and gene therapy

are some examples of emerging biomedical applications where biodegradable synthetic

polymers are being investigated.54-56

When considering degradable polymers for different

biomedical applications, it is important to ensure that they are bioresorbable. The

polymeric biomaterial is not only biocompatible, but upon degradation, the body can

effectively process and remove monomers, oligomers and byproducts.57

2.3.1.1 Poly(urethanes)

Poly(urethanes) (PURs) are an important class of polymers that have been used in a range

of high-performance materials including films, coatings, adhesives, fibres and elastomers.

Although they are biostable and have been extensively investigated for long-term medical

implants, their good biological properties, biocompatibility and synthetic versatility has

led to the development of biodegradable PURs. There are many different compounds that

can be applied to form PURs (by means of a simple polyaddition reaction) therefore

material properties are highly versatile.58

Typical preparations of polyurethanes involve a

polyester/polyether diol (soft segment), chain extender and a bulky diisocyanate (hard

segment). The multiblock structure is therefore responsible for giving PURs their

elastomeric properties.59

PURs were not considered to be thermoplastic until 1958, when

diphenylmethane-4, 4-diisocyanate (MDI) was incorporated as the bulky diisocyanate.60

However, in order to realize PURs as biodegradable polymers, common diisocyanates

such as MDI and toluene diisocyanate (TDI) cannot be employed due to their inherent

toxicity.9 Incorporation of lysine diioscyanate (LDI) or 1,4-diisocyanatobutane (BDI) can

alleviate these issues. Reacting LDI with polyester diols offers a range of applicable

properties (Figure 2.4 for structures).61

For example, peptide components in PURs allow

10

active moieties (ascorbic acid and glucose) to be introduced into the polymer thereby

promoting cell adhesion, viability and proliferation.62

Figure 2.4 – Structural representations of diisocyanates.

PURs possess excellent mechanical properties, including high tensile strength and

ultimate elongation due to their chemical structure, and because of this structure, can also

be processed via extrusion, injection molding and calendaring.63

The ease of

processability makes PURs attractive for use as injectable biodegradable polymers. This

has led to applications involving injectable hydrogels, which have been developed to

alleviate issues with current surgical techniques. In addition, an injectable LDI-based

polyurethane was developed (PolyNova®

) for orthopaedic applications because of its

good mechanical properties and fast self-setting as well as in vivo crosslinking ability.

Finally, porous scaffolds for TE of bone and cartilage have been proposed by usage of

PURs containing poly(caprolactone) PCL or poly(ethylene glycol) (PEG) segments due

to superior control over crystallinity (by controlling soft segment MW) and mechanical

properties.64,65

PUR-based scaffolds have also been investigated for both in vitro and in

vivo applications, such as analyzing cell density evolution [which was comparable to

biocompatible poly(lactic-co-glycolic acid)]66

and vascularization into dorsal skinfold

chambers of mice.67

Although the scaffolds did not appear to elicit any inflammatory

responses, acidic degradation of PURs autocatalyzes the degradation process and

byproduct production can lead to in vivo inflammatory responses.

11

2.3.1.2 Poly(ortho esters)

The development of poly(ortho esters) (POEs) produced biodegradable polymers higher

hydrophobicity, thereby avoiding issues with bulk degradation in drug delivery

applications.1,68

By imparting hydrolytically sensitive backbones, this would limit

degradation to very slow surface erosion in aqueous environments. There are four

families of POEs, each with varying syntheses to improve on shortcomings of the

preceding POEs (Figure 2.5).69

Figure 2.5 – Chemical structures of various POEs.

Because of versatility in their synthesis through incorporation of different diols, POEs

possess varying degradation rates, levels of pH sensitivity, and glass transition

temperatures.70

For drug release applications, the rate of release depends on the rate of

polymer hydrolysis. For example, the development of POE IV has led to the best control

in terms of release profile for various therapeutic molecules.71

With unprecedented

control over degradation rate as well as good biocompatibility evaluation of POE IV,

research has shifted towards using it as an injectable polymer for ocular applications,

treatment of periodontal diseases and estrus synchronization in sheep.72

2.3.1.3 Aliphatic Polyesters

Polyesters are defined as thermoplastic polymers with hydrolytically labile aliphatic ester

linkages throughout their backbone. This class of polymers is interesting due to its

diversity and synthetic versatility. These polymers can be prepared from a large range of

monomers through ring opening and condensation polymerization, resulting in materials

12

of very different properties. Furthermore, aliphatic polyesters are all biocompatible as

well as bioresorbable and FDA approved for a variety of applications.73,74

Although many

polymers have been studied for their potential applicability as biomaterials, polyesters,

such as poly(lactide) (PLA) and poly(caprolactone) (PCL) have recently been under

extensive investigation due to their exceptional biocompatibility as well as good

mechanical properties. In addition, both of these aliphatic polyesters have been shown to

generate bioresorbable metabolites during hydrolytic degradation, establishing their high

potential for replacing biostable polymers in time-limited applications. For example,

during hydrolytic degradation, PCL breaks down into different constituents that are

eventually eliminated from the body. It is first broken down into 6-hydroxyl caproic

acid, followed by Acetyl coenzyme A, which finally enters the citric acid cycle and is

excreted by the body (Scheme 2.1).75

PLA is also advantageous in terms of degradation

byproducts; they are metabolized into CO2 and water, or are excreted via the kidneys.76

These materials have therefore been heavily investigated for controlled drug release

systems and as orthopaedic implants.77-79

Scheme 2.1 – Hydrolysis of PCL to 6-hydroxylcaproic acid and acetyl coenzyme A

intermediates followed by elimination from the body through the citric acid cycle.

Both PCL and PLA are hydrophobic polymers that can be prepared via ring-opening

polymerization (ROP) using a variety of anionic, cationic and coordination catalysts, or

even via free-radical ROP.80

The mechanisms of initiation, coordination/insertion of

13

monomer, deactivation and chain transfer for PCL with stannous octoate can be

visualized in Scheme 2.2. First, an initiator bearing a hydroxyl-functionality (alcohol) is

added to react with stannous octoate producing a stannous alkoxide species and

ethylhexanoic acid (1). Next, further reaction with a second alcohol equivalent produces

the stannous dialkoxide initiator (2); adventitious water will serve as a catalyst

deactivator of either alkoxide initiator to a stannous alcohol derivative (3). Reaction of

the stannous dialkoxide initiator with monomer by coordination-insertion generates the

first actively propagating chain end, consisting of both the initiating alcohol fragment and

active propagating centre (4). Either further propagation will occur to grow the PCL

chain, or rapid intermolecular exchange of the stannous alkoxide moiety for a proton,

thereby establishing a rapid equilibrium between activated and deactivated chain ends

(5).

Scheme 2.2 – Mechanism of stannous octoate polymerization of PCL: 1/2 –

formation of stannous alkoxide initiator; 3 – deactivation of catalyst; 4 –

coordination/insertion of monomer; 5 – chain transfer of active polymerizing centre

to alcohol.

14

PCL, as well as two of PLA's stereochemical forms (poly-D-lactide and poly-L-lactide)

are semicrystalline materials, exhibiting glass transition temperatures (Tg) of -60°C, 55°C

and 60-65°C, and melting temperatures (Tm) of 59-64°C, 150-170°C and 175°C,

respectively.81-83

However, the third stereochemical form of PLA, poly-D,L-lactide

(PDLLA), is amorphous and therefore does not exhibit a Tm, rather, just a Tg of

approximately 55-60°C. In addition, poly-D-lactide (PDLA) is denoted as D due to its

ability to rotate a plane of polarized light to the right (clockwise, dextrorotatory);

conversely poly-L-lactide (PLLA) is denoted L as it rotates light to the left

(counterclockwise, levorotatory) (structures in Figure 2.6).9 Variations in transition

temperatures will impact their chemical and physical properties, providing insight toward

relative applicability in different biomedical applications.

Figure 2.6 – Chemical structures of PDLA (S-enantiomer), PLLA (R-enantiomer),

PDLLA (racemic mixture) and PCL.

Due to physical and chemical properties, these aliphatic polyesters have significantly

different degradation rates. PCL has been shown to break down exceedingly slowly, with

complete degradation requiring approximately a 3-4 year period.84

Polymeric devices

consisting of PCL first garnered interest for sustained drug release involving devices that

were to remain active for over 1 year, and as slowly degrading suture materials

(MaxonTM

).82

However, polyglycolides as well as polylactides became more popular for

drug delivery vehicles as they display the ability to completely release an encapsulated

drug over a few weeks and be fully resorbed in 2-4 months. Although PDLLA has been

shown to exhibit increased degradation rates, it typically starts to show mass loss and

fully erode within 12-16 months, whereas PLLA can take 2-5.6 years to degrade in

15

vivo.82,85

Degradation of semicrystalline polymers such as PCL occurs in two stages when

subjected to aqueous media. First, water diffuses into the amorphous regions, which are

less organized and allow water to penetrate more easily. Next, hydrolytic degradation

occurs from the edge to the centre of the crystalline domains, followed by intracellular

degradation if the MW is less than 3000 g/mol. This explains why PDLLA degrades

much faster as it lacks crystallinity.86-88

The second stage of degradation confirms that

PCL is resorbable, as polymer fragments are uptaken into phagosomes, whereby an

intracellular mechanism completes the degradation process. In terms of in vivo

degradation, PCL and PDLLA behave similarly.89

PCL’s excellent biocompatibility also makes it attractive for 3D porous scaffolds

in TE applications to direct the growth of cells and new bone at the site of

implantation.76,90

Moreover, its good rheological and viscoelastic properties render it easy

to manufacture and manipulate while providing overall structure and support.91,92

PCL’s

slow degradation and mechanical support are therefore excellent attributes for this

application. The slow degradation coupled with bioresorbability ensures ample time for

neo-bone/tissues to form at the site of implantation without complete fragmentation of the

biomaterial. Along with scaffolds, PCL has also been employed in various TE

applications including bone,93

cartilage,94,95

tendon and ligament,96

cardiovascular,97

blood vessel,98

skin99

and nerve.100

PCL is a highly versatile resorbable polymer and

although there are several FDA approved drug delivery and medical devices, an increase

in TE applications should emerge due to PCL-composite structures and their superior

mechanical and biocompatible properties.

PLA’s advantages also manifest from its biodegradable nature as well as high

strength and biocompatibility.101

PLA’s crystallinity depends on the ratio of D- and L-

enantiomers used. However, combinations with as little as 12% D-lactide result in the

amorphous PDLLA grade.102

PDLLA is commonly used in the food packaging sector due

to its ease of transformation (i.e. injection moulding and thermoforming), which also

makes it attractive for usage in resorbable plating, artificial cartilage or bone,

chemotherapeutic and pharmaceutical applications.103

These applications require faster

degradation, thereby portraying PDLLA’s advantage over its crystalline counterparts,

16

PDLA and PLLA. Because of its faster degradation and moderate strength, PDLLA is

preferably developed as a drug delivery vehicle or a scaffolding material for tissue

regeneration.9

By considering both PCL and PDLLA as possible bioresorbable polymers in this

study, it will provide consistency when comparing results. Although PDLLA has been

shown to degrade more quickly,77,89

degradation kinetics are based heavily upon MW, in

that higher MW polymers will take longer to degrade due to an increase in chain

length.104,105

Figure 2.7 illustrates the degradation profile for PCL homopolymers with a

linear/porous structure; linear PCL with an initial Mn of 30 000 g/mol elicited faster

initial and overall decrease in Mn (decreasing to 30% of its initial Mn after one year).106

Degradation is propagated due to hydrolytic degradation along the chain (or polymer

backbone), via surface or bulk degradation pathways. Surface degradation is more

predictable. The rate of hydrolytic cleavage, and therefore the production of oligomers

and monomers, which diffuse into the surroundings, is faster than the rate of water

infiltration into the polymer bulk. This does not cause a drastic change in MW, rather, an

overall thinning of the polymer.

Figure 2.7 – Molecular weight changes for a porous PCL structure (triangle) and a

linear PCL structure (square). Reproduced with permission from Taylor & Francis,

2007.

The current study involves the investigation of PCL and PDLLA grafted onto the IIR

backbone. IIR is known for its chemical and oxidative stability. Therefore PCL chains

will likely be localized within the IIR copolymer, thereby undergoing a reduced rate

degradation profile. Tethering of these polyesters to the IIR backbone will result in

17

compliance, yet stability and rigidity, perhaps making these copolymers suitable for a

variety of applications requiring good mechanical properties such as vascular prosthetics

or intervertebral disc replacement.

2.3.2 PIB and IIR Copolymers

2.3.2.1 Background

Poly(isobutylene) (PIB) is a synthetic elastomer, which yields many desirable properties

such as high elasticity, impermeability to gas and water, chemical stability and

biocompatibility. Because of these properties, PIB and its copolymers with small

percentages of isoprene (IP) (0.5-4mol%), commonly known as IIR, have been used in a

variety of commercial products such as the inner tubes of automobile tires, the bladders

of sporting equipment such as basketballs and soccer balls, lubricating oils, motor fuels,

sealants, and even as a primary component in chewing gum.107

Usage in these

applications is also possible due to its low level of unsaturation, which provides a route

for chemically crosslinking via sulfur-based curing. Crosslinking provides mechanical

improvements as well as abrasion resistance, thereby enhancing its physical properties

and bestowing suitability for different applications.108

PIB and IIR are attractive due to

their aforementioned properties and versatility, and their ability to be (co)polymerized via

cationic polymerization.

IIR was initially investigated by Gorianov and Butlerov (1870), as well as Otto

(1927); they found oily homopolymers of IB were successfully produced by usage of

boron trifluoride. By the 1930s, I.G. Farben Company of Germany fabricated high MW

PIBs, possessing rubber-like properties. The drawback of PIB was its inability to undergo

vulcanization or modification due to its fully saturated structure. Although uncurable,

homopolymers of PIB were commercialized from Badischer of Germany and Exxon

Chemical Company as PANOLand VISTANEX

, respectively. Additional research in

the 1930s conducted by W.J. Sparks and R.M. Thomas of Standard Oil and Development

Company (Exxon) allowed further development of IB into the first curable IB-based

elastomer, by incorporating small amounts of a diolefin, IP, into the molecule. However,

IIR was officially introduced and commercialized in 1942.109

In addition, halogen

18

derivatives such as chloro- and bromo-butyl were introduced in the 1960s as

commercially available products, which have greater variations in terms of vulcanization

and can be cured with other general elastomers. Butyl polymers are considered specialty

elastomers, eliciting the highest worldwide usage of all synthetic elastomers.

Versatility of PIB allows for the development of hybrid materials containing other

polymers, thereby imparting new properties to PIB for specific biomedical applications.

Demand for biomedical products is increasing in the western society due to the increasing

population of the elderly. Diversifying PIB’s usage toward the health sector is crucial to

satisfy the demand for products that enhance life quality and longevity. Although PIB is a

versatile polymer, approximately 80% of total PIB is directed toward the automobile

industry. The current clinical use of PIB-based copolymers in vascular stent coatings, as

well as its preclinical investigation in a number of other areas such as bone cements and

intervertebral disc replacements suggests that PIB is a highly biocompatible and

promising material for a range of biomedical applications.108

2.3.2.2 Synthesis of IIR

Commercial IIR grades such as poly(methylpropene-co-2-methyl-1,3 butadiene) or

poly(isobutylene-co-isoprene) are prepared by copolymerizing high purity IB and IP via

cationic polymerization at -100°C in methyl chloride. A schematic diagram of a typical

butyl plant can be found in Figure 2.8.110

Monomers and methyl chloride are purified via

flashing and stripping. Zinc or calcium stearate and antioxidants are added to prevent

agglomeration throughout the polymerization process. Post-reaction, the PIB product is

separated from the slurry, dried and processed. In addition, the reaction follows a generic

approach to provide living-like conditions, making use of conventional Lewis acid

initiation systems, but with the addition of a Lewis base. By employing Lewis acid

coinitiators (or activators) such as aluminum trichloride (AlCl3), alkylaluminum

dichloride and boron trifluoride (BF3) in methyl chloride or dimethyl sulphoxide (Lewis

base moderators), it modifies the interaction between the carbocation active centre and

counter-ion.111,112

Moreover, without this modified interaction, the counterion would be

too nucelophilic, causing the reactions to be terminated instantaneously.

19

Figure 2.8 – Commercialized IIR production: general IIR slurry polymerization

(Reproduced with permission from John Wiley & Sons, 1990).109

Polymerization of IB first involves the generation of a carbenium ion, which

occurs due to reaction of IB monomer with a Lewis acid catalyst such as AlCl3 (Scheme

2.3). Carbocation stability causes propagation to proceed mostly in successive head-to-

tail additions of monomer (either IB or IP) to the active centre.113

For IIR synthesis, IP

units typically enter the chain in a trans-1,4 configuration, as opposed to 1,2 and 3,4

modes of entry as evidenced by chemical analysis (Figure 2.9). The reaction is highly

exothermic therefore polymerization can be controlled by decreasing the temperature.

Methyl chloride is typically used as the reaction diluent with boiling liquid ethylene in

order to remove excess heat.114

Lastly, controlling factors such as temperature, solvent

polarity and the presence of counterions can also tune the rate of propagation.

Figure 2.9 – Isoprene unit enters chain predominantly in trans-1,4 configuration.

20

Propagation continues until either termination or chain transfer occurs. Termination

involves unimolecular rearrangement of the ion pair, causing the second last C-H bond to

break and release a proton, generating a terminal alkene bond. Termination can also

occur through alternate pathways such as formation of stable allylic carbenium ions, or

by carbocation reaction with nucleophilic species including amines or alcohols. Control

over termination allows for production of various MW IIR copolymers, capable of further

modification. Lastly, chain transfer to a monomer unit is the typical mechanism

governing polymerization termination (Scheme 2.3).115,116

The monomer effectively

abstracts a proton from the second last carbon (of the chain), resulting in a monomeric

carbocation. Both unimolecular rearrangement and chain transfer termination

mechanisms result in carbocation formation, thus propagating the growth of a new

polymer chain. However, chain transfer may also occur to solvent, impurities and

polymer chains resulting in branched polymers.

Scheme 2.3 – Cationic polymerization of IB governed by initiation, propagation and

termination.

21

2.3.2.3 Physical and Chemical Properties of IIR

The discovery of IIR was preceded by the desire to transform PIB into a rubbery

copolymer that allowed for low functionality, resulting from its low level of unsaturation.

As a result, low-modulus vulcanized networks that resist ozonolysis and oxidation can be

produced.117

In addition, because of its oxidative, enzymatic and hydrolytic resistance, it

is also biocompatible for long-term applications.118

Its biocompatibility is advantageous

for applications involving medical devices in vivo (vascular prosthetics, stents,

implantable devices, etc.), to replace materials such as PURs that may degrade, leading to

inflammatory and fibrotic reactions.119

The long, fully saturated PIB segments also

manifest physical properties such as low permeability to both gases and liquids, thermal

stability, weathering, chemical and moisture resistance as well as vibration damping.120

Very low permeability, making it advantageous for innerliner in tires, is attributed to the

efficient intermolecular packing and high density of the PIB segment. In terms of air

retention within tires, IIR demonstrated to be at least 8 times better than that of natural

rubber (Table 2.1).109

Table 2.1 – Air loss after automobile driving tests (Reproduced with permission

from John Wiley & Sons).109

Initial Conditions Air Pressure Loss (psi)

Inner Tube Original Pressure (psi) 1 Week 2 Weeks 3 Weeks

Natural Rubber 28 4.0 8.0 16.5

IIR 28 0.5 1.0 2.0

Moreover, PIB’s compact and symmetrical intermolecular packing minimizes its

intermolecular interactions, which is reflected by its viscoelastic properties.121

Having

two methyl side groups on every other chain carbon causes a delay in elastic response to

deformation. This can also be described as high hysteresis. Perfectly elastic materials do

not lose energy when a load is applied to them. However, viscoelastic materials do lose

some energy, dissipated as heat, resulting in slight permanent deformations after a given

stress is applied. This is also related to creep in that viscoelastic materials permanently

22

deform over prolonged periods of time due to the application of small stresses (strain rate

dependent on time). These characteristics, along with IIR’s high mechanical damping

make it viable for various automotive applications including suspension bumpers and fan

belts.121

In terms of chemical properties, IIR has a glass transition temperature of

approximately -70°C and is readily soluble in nonpolar solvents.122

Although typical IIR

exhibits IP contents ranging from approximately 0.5-4 mol%, it can be further increased

to 7 mol% to examine the impact of increased functionality and unsaturation. IP acts as a

strong chain transfer agent in the copolymerization of IB and IP, therefore conditions

must be tuned in order to afford IIR with high mol% of IP. Moreover, the reactivity ratios

of IB and IP monomers are similar, thus generating a randomly distributed copolymer.

Rapid and somewhat uncontrollable polymerization rates, chain transfer and termination

mechanisms contribute to IIR’s high polydispersity indices (PDI), ranging from 2-4.123

High PDIs indicate that IIR and copolymers have wide molecular mass distributions.

2.3.2.4 Modifications of IIR

Although IIR has attractive properties and application potential, chemically crosslinking

(vulcanizing) the elastomer improves abrasion resistance and mechanical properties.

Commonly employed vulcanization methods include accelerated sulfur vulcanization,

dioxime crosslinking or polymethylol-phenol resin curing. An example of sulfur-based

crosslinking is depicted in Scheme 2.4a. Sulfur compound varieties consisting of

thiurams, dithiocarbamates and thiazoles and concomitant temperatures of 160°C

generate crosslinked products from highly unsaturated IIRs. Scheme 2.4b describes

dioxime curing, where an oxidizing agent [O] oxidizes p-quinone dioxime, forming an

active crosslinking agent, which can rapidly vulcanize at room temperature (RT). Lastly,

Scheme 2.4c shows the general structure of a resin used for olefinic crosslinking. The

method is dependent on the R group reactivity.

23

Scheme 2.4 – Vulcanization of IIR: a) sulfur-based crosslinking; b) dioxime curing;

c) general structure of resin capable of vulcanization.

The discovery of IIR halogenations by Goodrich in the 1950s increased active

functionalities and therefore produced versatile IIR derivatives.124-126

While Goodrich

commercialized a brominated butyl (bromobutyl, BIIR) derivative in 1954, Hycar 2202,

Exxon researchers produced chlorobutyl (CIIR) and officially commercialized their

product by 1961. Halogenations consisted of “dark” reactions, performed with a solution

of IIR and elemental halogens (X) in hexane at approximately 40-65°C (Scheme 2.5).

The goal was to create a halogenated IIR compound with only 1 halogen atom per

isoprene unit (1:1 mole ratio of X to isoprene).127

BIIR and CIIR possessed the attractive

properties of IIR, as well as additional characteristics including enhanced cure properties

(broader vulcanization techniques) and covulcanization with other high-unsaturation

general-purpose elastomers.128,129

With these improvements, halobutyl tubeless tire

innerliners could be afforded. Typical addition reactions with halogen atoms result in

Zaitsev configuration; however, the reaction of IIR with either Cl2 or Br2 results

predominantly in substituted allylic halide structures, in this case, the exomethylene

isomer (Scheme 2.5). This product is the most kinetically favoured, as a result of steric

constraints from the dimethyl-substituted carbon. However, under thermal conditions,

this product can rearrange to the X-methyl isomer due to low strength exhibited by the

carbon-halogen single bond. Lastly, H-X elimination can occur producing conjugated

24

dienes, rendering the copolymer useless in terms of post-halogenation curing

techniques.130

Scheme 2.5 – IIR bromination followed by isomerization and HX elimination.

Chloro- and bromo- functionalities provide a greater variety in terms of curing

methods. For instance, zinc oxide (ZnO) can be used to cure both derivatives, followed

by simple extraction of Zn-X for post-purification.131

CIIR can also be vulcanized via bis-

alkylation or resin curing, producing an elastomer with increased heat resistance and

elastic modulus (E). However, BIIR has broader vulcanization versatility due the higher

reactivity of the C – Br bond allowing for straight sulfur cures, zinc-free cures and

peroxide cures in decreased reaction times. In addition, it possesses a higher affinity for

covulcanization with other unsaturated elastomers. Although chemical crosslinking of

halobutyl does impart increased stiffness, hardness, abrasion resistance and tensile

strength, harsh chemical-based crosslinking systems eliminate their potential for

biomedical applications. Therefore, milder conditions involving modification of the

backbone via installation of different chemical moieties, allows biologically acceptable

avenues for altering chemical and physical properties of IIR.

IIR has been functionalized with small molecules such as acids,132

esters,133,134

amines,135,136

and ethers.137

However, it is of interest to tune chemical and physical

properties via copolymerization or grafting, which allow a variety of polymers to be

directly conjugated to the IIR backbone. This should facilitate desirable chemical

characteristics to be expressed (originating from both polymers), resulting in interesting

25

hybrid materials. Over the past 20 years, PIB has been copolymerized to afford various

block copolymers. Poly(styrene-block-isobutylene-block-styrene), SIBS, which is a TPE,

has been prepared by a living cationic polymerization method of PIB and styrene

(Scheme 2.6).118,119

More specifically, SIBS was synthesized by first utilizing a

bifunctional initiator, 5-tert-butyl-1,3-bis(1-methoxy-l-methylethyl)-benzene (HDCE),

for the cationic polymerization of IB in methyl chloride/hexanes at -80ºC. After reaching

the desired MW of PIB, styrene was added and polymerized until termination via

methanol addition. SIBS has demonstrated desirable mechanical (ultimate tensile strength

and hardness) and stability properties in comparison to IIR owed to the styrene (glassy

and hard) blocks. In addition, utilizing trifunctional or arborescent initiators can afford

three-armed star SIBS block copolymers and PIB-based hyperbranched copolymers,

respectively.123,138

Scheme 2.6 – Reaction schematic elucidating SIBS production via bifunctional

HDCE initiatior.

In order to induce a broad range of characteristics, a multitude of polymers including

poly(methyl methacrylate) (PMMA),139-141

PLA142

and poly(ethylene oxide) (PEO),143,144

have also been used to fabricate linear and star-branched block copolymers with PIB.

Controlled polymerization conditions provide different molecular structures (di- and tri-

block, star and arborescent), thereby affording a vast library of PIB-based copolymers

possessing various properties.

26

IIR can also be functionalized via backbone grafting approach; directly attaching

polymers at the IP functionality affords IIR copolymers with interesting chemical,

physical and biological properties. In order to functionalize the unsaturated units of

rubber, epoxidation can afford moieties that can undergo further chemical modification

for various procedures. Zhang et. al. demonstrated epoxidation by means of

hydroperoxide compounds, in conjunction with molybdenum compounds to catalyze the

reaction.145,146

However, reaction conditions required temperatures of approximately

90°C and durations of 10 hours.

Scheme 2.7 – Synthesis of IIR-PEO graft copolymers.

In recent work, epoxidation of the isoprene functionality was performed at RT in the

presence of m-chloroperoxybenozic acid (mCPBA) for 1 hour, thus cleanly affording the

epoxidized product with over 99% conversion.147,148

This method not only required

milder conditions, but also eliminated the presence of transition metal catalysts and

prevented side-product formation. The epoxidized product was transformed via acidic

ring-opening catalysis in under 5 minutes to generate clean formation of a hydroxyl

(OH)-moiety. Interestingly, the product opposed Zaitsev’s rule, evidenced by formation

of the less-substituted alkene. By activation with p-nitrophenyl chloroformate (PNPC),

PEO consisting of varying MWs could be grafted to specifically tune PEO weight

percentage (wt%) in the IIR graft copolymers.147,148

Although hydroxyl-terminated PEO

could be used for grafting, PEO with amine-termini provided full conversion of the IP

units, due to very high coupling efficiencies (Scheme 2.7). Mild conditions coupled with

unprecedented control over graft copolymer fabrication and PEO content suggests the

grafting-to mechanism possesses greater industrial applicability.

27

Kawahara et. al. also investigated graft copolymers consisting of styrene and

deproteinized natural rubber (DPNR) in order to afford TPEs.149,150

In this study, graft

copolymerization was performed by employing tert-butyl hydroperoxide/

tetraethylenepentamine initiating systems with DPNR to achieve grafting efficiencies

between 70-95%. Although reasonable grafting efficiencies were afforded, control over

MW of the poly(styrene) (PS) portion was difficult due to deactivation and chain transfer.

In addition, the highest PS content attainable was 32wt%, which was attributed to the low

concentration of active site availability for grafting (5 x 10-4

mol/mol rubber compared to

2.2-7 mol% isoprene units for PEO-graft copolymers, Figure 2.10). The graph on the left

of Figure 2.10 shows that a critical amount of styrene monomer is needed to maintain

higher active grafting sites on the DPNR. Concentrations too low led to deactivation,

while higher concentrations resulted in chain transfer. Whereas the graph on the right

reiterates the importance of adding the correct concentration of monomer; lower feeds

ensure that styrene is grafted to rubber particles as opposed to forming PS homopolymers

or chain transfer products.150

Figure 2.10 – Monomer-concentration dependence on number of active sites and

styrene conversion (left); monomer-concentration dependence on MW of grafted PS

and MW of PS homopolymer (Biomaterials, 29, 2008, 448-460 © 2007 Elsevier Ltd.

with kind permission from Springer Science and Business Media).

28

2.4 Applications of IIR Materials

IIR’s high impermeability and flex fatigue impart it with appropriate properties for use in

tire fabrication, especially tire innerliners. The development of CIIR and BIIR butyl

derivatives, as mentioned, led to increased curing rates and versatility as covulcanization

with natural rubber (NR), butadiene rubber (BR) and styrene butadiene rubber (SBR)

became possible, resulting in tires with increased durability. Desirable properties

involving air and moisture impermeability and minimization of intercarcass pressure

(which could cause belt edge separation and adhesion failures), have caused CIIR and

BIIR to be used commercially.151

BIIR is advantageous for innerliner fabrication due to

increased adhesion to carcass compounds, better balance properties, lower density and

costs and flex-cracking resistance. Finally, various blends of NR and BR, NR and CIIR,

as well as SBR and BR have been used for tire black sidewall, white sidewall and tire

tread applications, respectively. In addition to tire applications, IIRs are also applied in

automotive hoses including coolant, fuel line and brake line hoses because of its

resistance to higher temperatures in under-the-hood applications.152,153

IIR's ability to

dampen vibrations also makes it suitable for incorporation in dynamic parts. IIR has even

been used in pharmaceutical applications such as IIR-stoppers and has been FDA

approved for chewing-gum due to its biological inertness.154

PIB-based linear copolymers exhibit excellent biocompatibility, thus broadening

their scope and applicability as potential biomaterials.108,155,156

For example, PIB-

copolymers are being developed as corneal shunts for the treatment of glaucoma,157

as

well as in synthetic aortic valves.158

PIB-PMMA composites have been shown to have

enhanced properties relative to commercial bone cements due to the incorporation of the

elastomeric PIB into the glassy PMMA material.139,159

However, limitations are related to

void formation throughout the material. These deficiencies led to inconsistencies in the

material itself, rendering it unsuitable for clinical use in bone cements. Multiarm PIB-

cyanoacrylate (CA) copolymers have been reported as promising materials for

intervertebral disk replacement due to the combination of CA chemistry and the

viscoelastic properties of PIB.160,161

Moreover, copolymers of PIB with hydrophilic

polymers such as poly(N,N-dimethylacrylamide) or PEO have been used to form

29

membranes that can encapsulate cells while allowing the exchange of oxygen, nutrients,

and secreted proteins such as insulin across the membrane.162

PEO incorporation into IIR

materials is also of interest due to PEO’s inherent nature to exhibit protein resistance.163

PIB-PEO copolymers could therefore be applied in applications where biofouling has

proven to be problematic.117

The level of protein adsorption resistance was investigated

via fluorescent confocal microscopy (Figure 2.11).148

Figure 2.11e and f, corresponding

to IIR-PEO graft copolymers with PEO incorporation of 24 and 34wt%, showed no

fluorescence, indicative of protein resistance. It is important to note that with increasing

PEO content, water solubility of IIR-PEO materials also increases. Structural integrity

could be compromised and may be problematic for potential biomedical applications.

Figure 2.11 – Fluorescence confocal microscopy images following adsorption of a

rhodamine-fibrinogen conjugate. PEO content: a) 2%, b) 4%, c) 6%, d) 12%, e)

24% and f) 34% (Reprinted with permission from Macromolecules 2011, 44, 6405.

Copyright (2011) American Chemical Society).

PIB-PS copolymers consist of soft segment elastomer and glassy PS domains,

resulting in TPEs (Figure 2.12).164

These TPEs showed excellent properties, similar to

that of IIR, whilst also exhibiting biostability and biocompatibility. With FDA approval,

these copolymers have been used for coating drug-eluting coronary stents (DES)

(Taxus® of Boston Scientific Co.).164

SIBS has also been investigated for use as a

30

corneal shunt, with results showing a decrease in inflammatory response when compared

to polydimethylsiloxane.165

These advances have led to the development of the MIDI-

tube, which has demonstrated high patency in rabbit test subjects for the treatment of

glaucoma. SIBS is also viable for the production of intraocular lenses166

as well as

trileaflet heart valve replacements.167,168

Its excellent mechanical properties as a TPE

makes it an attractive candidate for the aforementioned applications, however, creep

deformation may be problematic. Optimization of the polymer chemistry and resulting

properties is still critical for many applications. For example, when SIBS was explored as

a potential implant material in the urinary tract, significant attachment of uropathegenic

species such as E. coli 67 was observed, indicating that the surface properties of the

polymer were not ideal for this application.117

Furthermore, there have been reported

cases of stent coating delamination upon employment in vivo, indicating that the adhesion

of PIB-PS copolymers to the metal materials could be strengthened.169-171

Figure 2.12 – Cartoon representation of PIB-b-PS showing elastomeric

entanglements (PIB) and hard segments (PS) (Reprinted from Biomaterials, 29/4,

Pinchuk et. al., Medical applications of poly(styrene-block-isobutylene-block-

styrene) (“SIBS”) (448-460). Copyright (2008), with permission from Elsevier).

31

2.5 Evaluation of Biomaterials

2.5.1 Chemical Characterization

It is standard to characterize new polymeric materials by a range of techniques including

nuclear magnetic resonance (NMR) spectroscopy, size exclusion chromatography (SEC)

and fourier transform infrared spectroscopy (FTIR) in order to confirm their chemical

structures. NMR spectroscopy is a technique that allows one to obtain detailed

information on the chemical structures of molecules. While NMR spectroscopy can be

performed to probe any nucleus with a non-zero spin, one of the most commonly

investigated nuclei is the proton. In 1H NMR spectroscopy, each signal is characteristic of

a specific proton found in a given molecular structure. The precise resonance frequency is

affected by electron shielding, which is dependent on the chemical environment.

Therefore, this information is important in order to determine the chemical structure and

purity of materials. Moreover, the intensity of each peak can also allow for quantification

of a particular component in a sample. Specifically, by integrating copolymer peaks, the

determination of PCL or PDLLA incorporation for each copolymer can be successfully

calculated.

Gel permeation chromatography (GPC) is a type of size exclusion chromatography (SEC)

that separates analytes based on hydrodynamic volume, thereby providing information

about polymer chain length and in turn, MW. By utilizing packings comprising porous

beads in a column, separation is achieved. Smaller analytes enter the pores, therefore

increasing their retention time in the column, whereas larger analytes bypass the pores

thereby eluting much faster. In conventional SEC, a number of calibration standards of

known MW are run and their retention times are recorded in order to identify the

relationship between retention time and MW. By relating the retention time of an

unknown sample to the calibration curve, the unknown's MW can be determined. The

SEC results will describe different MWs: weight average molecular weight (Mw), number

average molecular weight (Mn), size average molecular weight (Mz) and the viscosity

molecular weight (Mν). Polymers can then be characterized in terms of PDI:

32

1)

Equation 1 – Definition of polydispersity index where Mn is the total weight of the

sample divided by the number of molecules (arithmetic mean) and Mw fairly

accounts for the contributions of different sized chains.

Fourier transform infrared spectroscopy (FTIR) involves the absorption of infrared light,

and is typically performed in order to identify the chemical functional groups present in a

sample. Since every sample is different, different bonds and molecules will lead to a very

specific spectrum, or ‘molecular fingerprint’. The analysis produces an interferogram (all

frequencies being measured at the same time), which undergoes a Fourier transformation,

providing a frequency spectrum. The peaks correspond to the frequencies of vibrations

between the atoms within chemical bonds. The variations in stretching are characteristic

of specific chemical functional groups such as carbonyls, hydroxyls, or alkenes. In

addition, peak intensity is directly related to the amount of a specific compound or

materials, allowing for quantitative analysis.

2.5.2 Physical Characterization

2.5.2.1 Water Contact Angle

The measurement of the water contact angle (WCA) of a surface can provide insight into

the hydrophobicity/hydrophilicty of a given surface, which can be crucial for various

applications. A given surface at a specific temperature and pressure will exhibit a unique

contact angle with a given liquid, which is affected by the surface energy and the

interfacial energy between the liquid and solid (dictated by cohesion and adhesion

forces).172

The critical surface energy (or critical surface tension) characterizes the

wettability of a material. Solids that possess high critical surface energies are more

wettable as opposed to solids with low critical surface energies, which are typically not

wettable by most liquids. WCA is measured by drop shape analysis, where the liquid

interface meets a solid surface as depicted in Figure 2.13.

33

Figure 2.13 – Contact angle measurement (θc) and interphase-energy between 3

phases (values in Young’s equation found below).

Although the results of this test will be used for qualitative analysis, the Young’s

equation can be used to find the surface energy of a solid:173

2)

Equation 2 – where ϴ = measured contact angle and γ is the surface tension of the

solid-gas (SG), solid-liquid (SG) and liquid-gas (LG) interface.

2.5.2.2 Atomic Force Microscopy

Due to advances in probe microscopies, techniques such as atomic force microscopy

(AFM) can provide a wealth of information on the surface topography, morphology, and

properties of materials.174

AFM is advantageous over other techniques (such as scanning

tunneling microscopy) as it can provide information concerning the nanomorphology of

the bulk polymer. There are various AFM techniques including contact AFM, contact

AFM in the light repulsive mode and lateral force AFM or friction force AFM. However,

in the following study, tapping mode will be employed, which provides short-range

forces that are still detectable, while minimizing the duration of tip-sample contact. The

forces of tip-contact in tapping mode AFM are approximately 0.1-1 nN (in comparison to

5-500 nN for contact mode).175

These low forces in combination with intermittent contact

result in low lateral drag forces, thereby reducing damage of soft polymers.176

This is

particularly important, as the IIR-copolymers are relatively soft. Low forces will help

minimize artifacts that could be produced when the tip contacts the material surface. In

addition, because tapping is conducted normal to the surface, it decreases capillary and

adhesion forces, providing better resolution as compared to static scanning AFM.

34

Tapping mode is conducted through use of a cantilever and tip (probe). Forces

occur between the cantilever and the surface, which are determined by the resulting

deflection of the cantilever beam. Moreover, a piezoelectric crystal causes the cantilever

to oscillate near its resonance frequency (average of 300 kHz). The amplitude of

oscillation decreases as the cantilever nears the surface, which is a result of energy loss

manifested by Van der Waals and electrostatic forces. It is because of this decrease in

amplitude that the surface features can be measured and identified. However, an

electronic servo must be used in order to maintain the cantilever oscillation amplitude by

means of a feedback mechanism (adjusts the tip-sample separation distance). The

software automatically sets the frequency to maintain the lowest possible level of force

on the sample. In doing so, the oscillation amplitude can be accurately measured by the

detector and input into the controller electronics of the instrument. Both the topographical

and phase images (detections) are produced simultaneously by converting the force of the

tip contacting the surface into an analyzable image. These images can reveal

subnanometer surface chemical resolution due to mechanical differences amongst various

domains.177

To obtain phase information, a phase shift is detected between the driving

and actual tip response oscillation signals. When copolymers or blends are studied via

AFM, one component displays lower surface energy and therefore typically dominates

the top few angstroms of the surface.178

2.5.2.3 Scanning Electron Microscopy

Scanning electron microscopy (SEM) is a commonly used imaging technique for the

determination of surface morphology, three-dimensional (3D) structures, and can also

provide information on chemical composition. The scanning electron microscope utilizes

a focused beam of accelerated electrons, thus generating a variety of signals at the surface

of the test material. Signals that produce the 3D images consist of secondary and

backscattered electrons. Secondary electrons show morphology and topography, while

backscattered electrons show contrasts in composition of multiphase systems. Finally, the

electron beam scans in a raster scan pattern (image capture and reconstruction),

combining the beam’s position with the detected signal to produce the final image.

35

2.5.2.4 Thermal Properties

Differential scanning calorimetry (DSC) is a thermal analysis technique used to measure

heat capacity changes of a given material with respect to temperature for detection of

glass (Tg) and melt transitions (Tm), crystallization, mesomorphic transition temperatures

as well as phase changes and curing/crosslinking. Understanding thermal properties

provides insight relating to the amorphous, semicrystalline or crystalline properties of a

given material. Amorphous materials are those that are completely non-crystalline.

Polymer glasses and rubbers make up the majority of such materials.113

When amorphous

materials are in the solid state, they are considered to be frozen polymer liquids that are

inherently hard and brittle (at low temperatures). However, as the temperature increases

and reaches the Tg, the polymer transforms to a rubbery material, gaining the ability to

flow (non-frozen).179

The glass transition is an important temperature as it dictates

polymer properties. Temperatures lower than Tg result in a dramatic increase in stiffness

relative to higher temperatures, as well as changes to the physical properties (heat

capacity and thermal expansion coefficient) of an amorphous polymer. These physical

properties will vary at temperatures surpassing the Tg. Polymers possess different Tgs,

which are dictated by their chemical structure, molar mass, branching, chain flexibility,

copolymerization and molecular architecture.

Crystalline or semicrystalline polymers possess the ability to crystallize

(thermodynamically favoured to reduce Gibbs free energy, G, or kinetically when cooled

quickly) when cooled below the melting point of the crystalline phase. Many factors

affect the rate and extent of crystallization for a particular polymer such as rate of

cooling, orientation in the melt, specific melt temperature, tacticity, chain branching and

molar mass.180

For crystalline polymers, atoms are covalently bonded into tightly packed,

unidirectional macromolecular arrays. Unit cells are made up of repeating segments of

polymer chains, with several segments of varying complexity in each unit. The polymer

chains lie in one particular direction resulting in relatively weak bonding between the

molecules, resulting in anisotropic physical properties. However, semicrystalline

polymers are of particular importance; melt-crystallized polymers are never completely

crystalline due to a large number of chain entanglements, thereby making it near

36

impossible to form an entirely crystalline polymer.181

Semicrystalline polymers contain

both crystalline and amorphous components due to irregularities in crystal formation.

Lamellar crystals are separated from each other by layers of amorphous phase, with

lamellae thickness dictated by interfacial energies, Tg and Tm, under-cooling and

segmental diffusivity.182

As with amorphous polymers, the transition temperatures of

semicrystalline polymers are affected by chemical structure (stiffness, polar groups, side

groups), molar mass and branching, as well as copolymer structure. In a copolymer,

typically, only one of the polymers is crystallizable, causing the melt temperature to be

impacted by incorporation into the copolymer. Main-chain stiffness is the predominant

factor in determining Tm and Tg, resulting in a correlation that affects both of these

properties. Typically, the value of Tg (K) is between 0.5Tm and 0.8Tg.183

Control over

these transitions is particularly evident in copolymers, thereby making these materials

useful for influencing chemical and physical properties.

Thermal characterization can even allow one to make inferences regarding

mechanical properties (Figure 2.14).184

Rubbery materials are those that exhibit only

glass transition temperatures due to their long polymeric chains and high degrees of

flexibility/mobility, allowing them to undergo large deformations. In other words, the

response of rubber is intramolecular; externally applied forces are transmitted to the long

chains through linkages at their outer peripheries changing their conformations, thus

causing each chain to act like an individual spring.185

Long chains tend to alter their

configuration rapidly, thus allowing typical rubbers to be stretched up to 10 times their

original length. Removal of external forces leads to rapid restoration of original

dimensions. However, when crystalline and glassy materials are subjected to external

forces, deformations that cause two neighbouring atoms to be altered by more than a few

angstroms will lead to unrecoverable deformations.

37

Figure 2.14 – Temperature dependence of stiffness of typical thermoplastic

elastomers (Reprinted from Handbook of Thermoplastic Elastomers, 1st Edition,

Drobny, Jiri George, Introduction (1-7). Copyright (2007), with permission from

Elsevier).58

Materials of particular interest are TPEs, which are composed of a hard

semicrystalline phase, and a soft amorphous phase. Phases are most commonly bonded

chemically through block/graft copolymerization.7 Performing DSC is therefore

important to first determine Tgs and Tms of polymers that may be used to make up a

particular TPE. Each individual polymer typically retains most of its characteristics, with

slight variations. The importance of the Tg and Tm relates to the physical variations of the

elastomer phase. As shown in Figure 2.14, temperatures below the Tg of the elastomeric

phase cause both phases to be hard, resulting in a stiff and brittle material. However,

above the Tg of the amorphous phase, softening occurs, producing an elastic material with

rigidity supplied by the hard phase, causing TPEs to behave similarly to vulcanized

rubber. With increasing temperature, the hard phase will eventually melt, producing a

viscous fluid. Performing DSC is pertinent to understanding the ideal service temperature

38

for TPEs which will be above the Tg of the rubbery phase, but below the Tm of the hard

phase. Overall, TPEs are attractive when compared to vulcanized rubbers due to simpler

and milder processing methods.58

In this thesis, investigating IIR-polyester graft copolymers will provide materials

that can be potentially processed as TPEs. Moreover, these graft copolymers may belong

to an important family of TPEs, polyester elastomers (PEE). These TPEs share

similarities to polyurethane and poly(amide) TPEs.186

Polyester elastomers, like other

TPEs, resist deformation due to the presence of microcrystallites formed by partial

crystallization of hard segments, which therefore function as physical crosslinks.

Processing temperatures allow these crystallites to melt, yielding a polymer melt; such

polymers can be shaped by moulding and retain their shape upon cooling as the hard

segments recrystallize.184

This aspect could be advantageous for various applications

involving injectable thermosets or the fabrication of specifically shaped implants. PEEs

are also highly versatile. Varying the ratio of hard to soft segment can result in materials

ranging from soft elastomers to relatively hard elastoplastics.

2.5.2.5 Mechanical Properties

Tensile testing is considered to be the most fundamental type of mechanical testing that

can be performed on a specific material. This testing mechanism provides useful

information on the material's ultimate tensile strength (UTS), Young’s Modulus (modulus

of elasticity, E), yield strength and elongation at break (εb). Identifying this information is

particularly useful for research and development, engineering design and quality control

and specification. The instrument itself utilizes a pair of self-aligning grips where the

sample is placed and secured, ensuring that the sample is aligned with the direction of

pull and to avoid possible slippage (Figure 2.15). Samples are typically stretched

uniaxally until failure by gradually increasing the tensile load (breakage). Tensile testing

machines therefore elongate the specimen at a constant rate, continuously and

simultaneously measuring the instantaneous applied load and resulting elongations. The

load-deformation characteristics are dependent on the specimen size.

39

Figure 2.15 – Instron (3300 series) tensile testing instrument.

Load and elongation are normalized to parameters of engineering stress (σ) and

strain (ε). Stress occurs when a force (F) is applied normal to the face of an element

(Figure 2.16a). The force transmits through the element and is balanced by an equal force

on the opposite side, establishing equilibrium. Strain is the response of materials to an

applied stress, causing the given material to stretch from its original length, Lo, to a final

length of L (δL = L – Lo). Upon application of a given stress and strain response, elastic

deformation occurs, which is described by Hooke’s law, stating that stress is proportional

to strain. Hooke’s law allows the Young’s modulus, E, to be defined for a material using

simple uniaxial extension given by:

3)

Equation 3 – Young’s modulus determination for a material at a given strain and

stress.

However, this relationship only occurs in the linear portion of a particular stress versus

strain trace resulting from a tensile stress. Within the elastic limit, a material will return

to its initial shape upon removal of the applied stress. Once the limit is surpassed,

permanent deformation will result. The modulus, E, can be considered to be a material’s

stiffness or resistance to elastic deformation. Higher moduli are measured in stiffer

40

materials.187

Certain materials, particularly polymers, do not obey Hooke’s law. When a

uniaxial stress is applied, although there is stretching and elongation in the direction of

the stress, this elongation causes constrictions (strains) to occur in the lateral directions

(εt), perpendicular to the applied stress. In addition to axial stress, there exists stress

parallel to the face of an object resulting in shear strain (Figure 2.16b) as well as equal

tensile/compressive forces to all six faces of a cubic element, resulting in hydrostatic

pressure stress (volume strain, or dilatation, Figure 2.16c). Therefore, as with E, there are

the shear modulus (G) and the bulk modulus (K), showing linearity between shear stress

and shear strain and hydrostatic pressure and dilatation, respectively:

4)

Equation 4 – Linear relationship relating the shear strain, γ to the shear stress, τ.

5)

Equation 5 – Linear relationship showing proportionality between the dilatation, Δ

and pressure, p.

If the material behaves isotropically (strains are equal in lateral directions), a parameter

termed the Poisson’s ratio is defined as the ratio of lateral and axial strains:

6)

Equation 6 – Definition of Poisson’s ratio where εt is the transverse strain and ε is

the axial strain.

When an element is stretched axially in one direction, the element contracts in the lateral

or transverse directions, resulting in a positive value for Poisson’s ratio, typically in the

range between 0.25 and 0.35 (for most materials).188

Therefore the definition of Poisson’s

ratio (Equation 6) successfully relates all three moduli (E, G and K) to one another for an

isotropic material as:

41

;

7)

Equation 7 – Relation between Young’s modulus (E), shear modulus (G) and bulk

modulus (K).

Figure 2.16 – Definitions of uniaxial stress, strain and elastic deformation.

(Reprinted from Nanomaterials, Nanotechnologies and Design: An Introduction for

Engineers and Architects, 1st Edition, Ashby, Michael F., Chapter 4 – Material

Classes, Structure and properties. Copyright (2009), with permission from

Elsevier).189

As shown in Figure 2.17, yield properties/ductility is measured using tensile tests

by taking the material to failure. The yield strength, σy, depicts the stress at which the

stress-strain curve (in the linear elastic regime) for axial loading deviates by a strain of

1% (for polymers). However, the behaviour beyond yield depends on the temperature

relative to the polymer’s characteristic Tg. Below the Tg, most polymers are brittle and

42

exhibit brittle fracture. At temperatures approaching the Tg, plasticity is possible and once

the Tg is reached, cold drawing is achieved. This is a large plastic (permanent

deformation) extension at constant stress during which molecules are pulled into

alignment with the direction of straining, followed by hardening and fracture. At higher

temperatures, thermoplastics become viscous and can therefore be moulded. Finally,

plastic strain, εpl, is the permanent strain that results from plasticity, defined as the total

strain, εtot, minus the recoverable, elastic part:

8)

Equation 8 – Definition of plastic strain.

Figure 2.17 – Stress strain curve for a polymer. Definitions of uniaxial stress, strain

and elastic deformation. (Reprinted from Nanomaterials, Nanotechnologies and

Design: An Introduction for Engineers and Architects, 1st Edition, Ashby, Michael

F., Chapter 4 – Material Classes, Structure and properties. Copyright (2009), with

permission from Elsevier).189

Polymers obey Hooke’s Law at low strains, therefore allowing the calculation of the

Young’s modulus by using appropriate software. However, with many elastomers and

semicrystalline polymers, the linear portion is difficult to define. Because of this, moduli

may be determined by tangent or secant methods. The tangent is the value of E at any

point in a curve, whereas in the secant method, the curve is bisected, and E (the slope) is

determined for the bisecting line.

43

Elastomers typically have very low Young’s moduli, in the approximate range of

0.5-1 MPa, whereas semicrystalline polymers exhibit higher values for E and UTS.

Figure 2.18a shows the broad range of moduli for a variety of material classes with large

differences in density. Polymers and elastomers possess densities lower than metals or

ceramics, as well as moduli below 10 GPa (elastomers specifically in the range of 1-10

MPa). Figure 2.18b depicts the yield strain of various materials. The yield strain (σy/E) is

the strain at which a material deviates from the elastic linear regime. It is important to

note that elastomers, due to their extremely low moduli, display yield strains in the range

of 1 to 10, the highest of all materials. This is important as a larger yield strain

corresponds to greater resistance to brittle fracture, which is important for various

biomedical applications.

44

Figure 2.18 – Material property charts showing: a) Young’s modulus and its

relation to material density; b) Young’s modulus and material strength to defined

the yield strain (σy/E), where a material no longer behaves elastically. (Reprinted

from Nanomaterials, Nanotechnologies and Design: An Introduction for Engineers

and Architects, 1st Edition, Ashby, Michael F., Chapter 4 – Material Classes,

Structure and properties. Copyright (2009), with permission from Elsevier).189

Elastomers are unique because although they are not physically stiff, they are able

to stretch out of the linear regime without resulting in permanent deformation. In contrast

to elastomers, semicrystalline polymers are considered to be a mix of polymer crystals,

randomly distributed throughout an amorphous matrix. If the amorphous phase is above

its glass transition temperature, the polymer will be less brittle, therefore exhibiting a

modulus of approximately 50-100 GPa.113

An increase in crystallinity dramatically

45

affects mechanical behaviour of a given polymer. At lower degrees of crystallinity, such

as those achieved by grafting PCL/PDLLA onto the IIR backbone, the crystalline

domains throughout the amorphous rubber should behave as crosslinks, producing

stiffness by increasing crosslink density. However, because PCL/PDLLA homopolymers

have higher degrees of crystallinity than the copolymers, the resulting moduli will be a

combination affected by both the amorphous and crystalline regions.

2.5.3 Biological Characterization

2.5.3.1 Cell-Material Interaction

Performing adequate assessment in terms of biological risks is inherently important when

fabricating potential biomaterials. In order to properly assess biological safety, first, the

toxicology of chemical constituents used in a potential biomaterial need to be

scrutinized.190

IIR, PCL and PDLLA are considered to be biocompatible materials;

however, modes of preparation (various chemicals/solvents) of graft copolymers could

potentially affect their biological, material and cellular responses. Since these materials

are intended for use in a biological system, it is important to assess if they are

biologically compatible with tissues, to minimize risk for a given patient.191

In order for a

device to be considered biocompatible, it must be able perform with an appropriate host

response in a specific application.192

Measuring biocompatibility is problematic due to

the breadth of applications, which involve the interaction of different materials within

different biological systems. However, in order to assess biocompatibility, material

acceptance in vivo should be used to evaluate potential applications.193

Furthermore, the

International Organization for Standardization (ISO) has prepared a guideline document:

Biological Testing of Medical Devices—Part 1: Guidance on Selection of Tests’ (ISO

10933-1) to provide insight into testing methods that should be employed. Depending on

whether the material will be in contact with the cardiovascular system, implanted, blood-

interfacing, skin/bone-contacting, or other will impact the relevancy of specific testing

methods.

46

Figure 2.19 – PCL and PCL-Col (collagen) materials: confocal scanning laser

microscopies showing differences in cell proliferation on different surfaces

(Reprinted from Biomaterials, 25/11, Cheng and Teoh, Surface modification of ultra

thin poly (ε-caprolactone) films using acrylic acid and collagen (1991-2001).

Copyright (2003), with permission from Elsevier).

It is fundamental to explore cytotoxicities of materials, as well as interactions of

biosystems with materials at molecular levels. Investigating the interaction between

mammalian cells and biomaterials via spreadability, adhesion and proliferation properties

is particularly crucial.194

Cells tend to communicate with their surroundings by means of

cell-surface interactions involving the formation of focal adhesions as well as the

clustering of integrin receptors.195

Physical characterization techniques that analyze

factors such as wettability, chemical composition196

and mechanical197

and

topographical198

properties can affect cell adhesion and therefore proliferation.

Investigating cell spreading on surfaces is important as anchorage-dependent cells need

to strongly adhere in order to differentiate and transfer signals and maintain cell

homeostasis. Through investigation of human myoblasts cultured on PCL films, adherent

cells did not spread and were consequently washed from the film. However, through

surface modification to incorporate collagen (Figure 2.19), cells spread in all directions

and experienced an increase in proliferation rate, thereby establishing a relationship

between adhesion and cell survival capabilities.199

Additionally, cells that maintain a

circular shape typically do not display actin stress fibres within the cytoplasm, which are

47

critical when determining cellular health. Shin et. al. conducted studies by seeding

Human Mesenchymal Stem Cells (hMSC) onto surfaces of Poly(L-lactide-co-E-

caprolactone) (PLCL) and compared them to surfaces of PLCL conjugated with acrylic

acid (AAc) and gelatin (Figure 2.20).195

Cells on PLCL did not exhibit any spreading, but

through surface modification, the morphology of the cells on the gelatin-AAc-PLCL

displayed cells polygonally elongated in shape.

Figure 2.20 – Morphologies of hMSCs cultured on various substrates: B) glass

control; C) PLCL; D) AAc-PLCL and E) gelatin-AAc-PLCL. Scale bar = 200 μm

(Reprinted with permission from Shin, Y. M.; Kim, K.-S.; Lim, Y. M.; Nho, Y. C.;

Shin, H. Biomacromolecules 2008, 9, 1772. Copyright 2008 American Chemical

Society).

C2C12, which is a murine myoblast cell line, possesses advantages including its

ability to rapidly differentiate, excellent fusion and production of characteristic muscle

fibre proteins.200

Myoblasts, which are representative of either a muscle cell or fibre, are

interesting to examine due to their end-to-end fusion configuration in vitro, producing

morphologies and spatial arrangements in an elongated and predictable manner.201

Dugan

et. al. fabricated cellulose nanowhiskers (CNW) to understand the proliferation of C2C12

murine myoblasts, as well as the differentiation and fusion to form myotubes (muscle

fibres).200

Focal adhesions as well as the F-actin in the cytoskeleton were stained and

imaged with confocal microscopy to determine if the morphology of CNW surfaces

48

affected cellular growth and attachment (Figure 2.21). The top line images show

differences in morphology in that myoblasts on the glass coverslip were less spread when

compared to CNWs (prepared at 500 and 6000 rpm); CNWs also exhibited vinculin (light

specks), indicating adhesivity to surfaces. Morphology differences were attributed to

surface roughness; roughness of various materials can affect cellular growth response.202

Although stress fibres appeared on all surfaces, variations in focal adhesion orientation

became apparent after 12 hours of incubation.

Figure 2.21 – (a-f) Myoblasts stained for vinculin (light speckles), F-actin (outer

periphery) and nuclei, scale bar = 50 μm. (a) 4 h after seeding on glass coverslip

control, (b) 4 h after seeding on C500 surface, (c) 4 h after seeding on C6000 surface,

(d) 12 h after seeding on glass coverslip control, (e) 12 h after seeding on C500

surface, and (f) 12 h after seeding on C6000 surface (Reprinted with permission

from Dugan, J. M.; Gough, J. E.; Eichhorn, S. J. Biomacromolecules 2010, 11, 2498,

Copyright 2010 American Chemical Society).

Finally, protein adsorption is a spontaneous occurrence with a high level of

importance in biomaterials and biomedical science.203

Although biofouling is

nondesirable, the adsorption of cell-adhesive proteins (such as vinculin and F-actin) may

be needed to some extent depending on the intended application. This adsorption can be

controlled in various ways such as modifying surfaces with polymer brushes (polymer

chains attached to a surface),204,205

resulting in stimuli-responsive signals such as

49

temperature, pH and light. Polymer brushes may be modified by attaching polymers that

are known to adsorb proteins with polymers that are protein-repellent to finely tune

overall adsorption.206,207

Moreover, characteristic protein adsorption exhibited by

polymers is dictated by their specific chemical structure. For example, a predominantly

studied protein-repellent polymer is poly(ethylene oxide) (PEO). Its protein-resistant

properties are owed to steric repulsion, causing the polymer to prevent proteins from

reaching the substrate surface to adsorb.208,209

Figure 2.22 – Confocal microscopy images of C2C12 cells adhered to control and

copolymer surfaces: a) glass (control); b) IIR (control); c) 18wt% PEO ; d) 32wt%

PEO; e) 65 wt% PEO; f) 83 wt% PEO. Nuclei (dark inner portion) and F-actin

fibres (lighter periphery) with image area = 0.22 x 0.22 mm (Reprinted with

permission from John Wiley & Sons, 2013).

Factors including grafting density, length and conformation of PEO chains can also

influence its surface resistance to proteins.210,211

In particular, IIR-PEO graft copolymers

have been fabricated for applications involving increased hydrophilicity and therefore

emulsifying ability,212

as well as varying PEO incorporation to confer resistance of the

surface to proteins.147,148,213

Proper growth and proliferation of cells on surfaces is

believed to be dictated by the ability of proteins to properly adsorb to surface substrates.

Recently, Karamdoust et al. showed that by increasing PEO incorporation into IIR-PEO

50

graft copolymers, reaching a critical PEO content (34 wt%, Figure 2.22d) caused a

dramatic decrease in protein adsorption, evidenced by a decrease in cell adhesion.214

51

2.6 Thesis Objectives

The goal of this thesis is to synthesize IIR-polyester graft copolymers and to study their

chemical, physical, mechanical and biological properties in order to gauge their potential

as biomaterials. An advantage of this approach is that rather than using chemical

crosslinking methods that are incompatible with biomedical applications, the thermal and

mechanical properties of the polyester components may impart enhanced strength to IIR

and other properties that are desirable for specific applications. In addition, the slow but

eventual degradability of the polyester component may assist in the gradual

environmental degradation of IIR materials, which are otherwise broken down only

extremely slowly in nature.

The interest in IIR-polyester graft copolymers as potential biomaterials stems

from the widespread interest in IIR for various biomedical applications. For example,

current bone cement applications involve usage of poly(methyl methacrylate), which is

inherently brittle.139,140

To overcome this issue, toughening of bone cements can be

accomplished with IIR-graft copolymer synthesis to impart impact- and fatigue-resistance

properties. Similarly, PIB-CA materials are intended for intervertebral disc replacement,

however, enzymatic attack on CA moieties causes the release of toxic byproducts. PURs

are employed for usage in vascular grafts due to their elastomeric nature, but chemical

and mechanical deficiencies led to inflammatory (and therefore occlusion) and crack

manifestation, respectively.118

Therefore, incorporation of bioresorbable and biodegradable polyesters may

eliminate issues associated with toxicity as well as the resistance of synthetic polymers to

degradation for time-limited applications, including sutures and bone fixation

devices.57,215

Utilization of stable polymers can lead to undesirable inflammatory

responses, as biological systems recognize them as foreign substances. However,

degradable polymers offer advantages for therapeutic applications, specifically in

medicine, surgery or drug delivery.9 For time-limited applications involving degradation,

bioresorbability is important as it describes materials with non-toxic byproducts that can

be eliminated from the body through metabolic pathways.76

52

In order to synthesize the IIR-polyester graft copolymers, the aim will be the

development of a simple synthetic method. This will make the final materials more

attractive on an industrial as well as biomedical level. The exploration of both "grafting

from" and "grafting to" methods will be described. Using the more successful "grafting

to" strategy, a small library of graft copolymers is produced to provide insight concerning

how varying weight percentages of polyester with respect to IIR affect chemical and

physical properties. These are characterized chemically by a variety of techniques

including NMR and IR spectroscopy, and SEC.

From differences in the chemical compositions of the materials, it is demonstrated

that differences in physical properties arise. These are assessed and compared using

techniques such as AFM, SEM, water contact measurements, and DSC. Tensile testing is

used to elucidate changes in the mechanical properties of the materials as a function of

their composition and degradation studies are performed to investigate their

degradabilities. Lastly, although the literature suggests that IIR,216

PCL82

and PDLLA88

are all biocompatible materials, the toxicities and cell adhering/proliferation properties of

the new graft copolymers are studied in this thesis. Combined, this data set provides a

basis of important information concerning the properties of these materials for further use

in specific applications.

53

Chapter 3

3 Results and Discussion

3.1 Synthesis and Chemical Characterization of IIR-PCL Copolymers

3.1.1 ROP of ε-caprolactone from the IIR Backbone

IIR-polyester graft copolymers via a "grafting from" approach was initially explored. The

goal was to perform a ROP of ε-caprolactone from the hydroxyl-moieties of the IIR

derivative 3.3 (Scheme 3.1). First 3.3 was prepared as previously reported,147

via

epoxidation of commercially available IIR (RB-402) (3.1) to provide the epoxidized IIR

derivative 3.2, followed by ring opening under acidic conditions to obtain 3.3. A library

of graft copolymers were to be synthesized with varying weight percentages of PCL, as

depicted in Table 3.1.

Scheme 3.1 – Epoxidation followed by hydroxylation of IIR with 2.2mol% IP and

subsequent ROP of ε-caprolactone from IIR backbone.

54

Table 3.1 – Varying conditions to afford IIR and PCL graft copolymers by ROP of

ε-caprolactone from –OH moiety on IIR backbone.

Target PCL wt% of

Copolymer

Time Catalyst Amount MSA or

Sn (Oct)2 (eq/OH Butyl)

75 6h - overnight 1-3

50 2.5h -

overnight

1-2

25 3h - overnight 1-1.5

10 4h - overnight 1-1.2

The initial preparation of graft copolymers via polymerization from the IIR backbone

hydroxyl moieties appeared successful. ε-caprolactone monomer polymerized, as

evidenced by the triplet corresponding to PCL at 4.06 ppm (-CH2 adjacent to oxygen),

and IB peaks were visible at 1.12 and 1.40 ppm (Figure 3.1b). Moreover, 1H NMR

spectroscopy also showed that the target PCL weight percentages (10-75 wt%, as per

Table 3.1) were obtained, based on integration of the PCL triplet at 4.06 ppm with respect

to the IB singlet at 1.40 ppm. However, it was later found that as the graft copolymers

were initially precipitated into methanol, this solvent also caused all free PCL oligomers

to precipitate. When the graft copolymers were precipitated in acetone, the free PCL

remained soluble, leaving the purified graft copolymer. Acetone and similar solvents (2-

butanone) solubilize PCL with MWs of 20 000 g/mol or lower, thereby providing a

means to elimination of free homopolymer. Upon reprecipitation of the copolymers into

acetone, it became apparent that the graft copolymer products mainly contained free PCL

trapped within the rubber. Thus, ε-caprolactone was preferentially undergoing

homopolymerization, likely from adventitious water impurities, rather than from the

hydroxyl functionalities along the IIR backbone. Therefore subsequent precipitations of

graft copolymers resulted in a significant decrease in PCL content. This aspect was

confirmed by taking additional NMR spectra of the copolymer material following

acetone precipitation (Figure 3.1c). By integrating the aforementioned peaks associated

with PCL and IB, a large decrease in PCL relative to IIR was revealed, in comparison to

the initial spectrum (Figure 3.1b).

55

Figure 3.1 – NMR spectra showing: a) hydroxylated IIR; b) 50 wt% IIR-PCL graft

copolymer following methanol precipitation (label k denotes terminal methylene

PCL); c) second precipitation in acetone of the same polymer from b), confirming a

decrease in PCL content to 16 wt% (label k denotes terminal PCL methylene).

Aside from 1H NMR analysis, issues with the ROP of ε-caprolactone from hydroxyl

moieties along the IIR-backbone were also confirmed by SEC and DSC traces. The

aforementioned homopolymerization of ε-caprolactone coupled with precipitations into

methanol, caused SEC traces to display significant side peaks (at increased retention

times), representative of a lower MW homopolymer, PCL (Figure 3.2, 50 wt% PCL). In

addition, DSC traces also depicted two melting temperatures (Appendix F). These

different Tms signified the existence of two different species: PCL homopolymer and IIR-

PCL graft copolymers. The lower Tm (43°C) belonged to the graft copolymer, as covalent

attachment to IIR has been shown to decrease the melting temperature of semicrystalline

polymers.147

The higher Tm value of 53°C was attributed to PCL homopolymer. Free PCL

56

chains throughout the rubber matrix would provide larger crystalline domains, thereby

showing an increased value of Tm. However, reprecipitation of graft copolymers into

acetone markedly increased copolymer purity as free PCL chains were removed,

producing copolymers with PCL successfully grafted to the IP functionality. This was

confirmed by obtaining a second SEC (Figure 3.2, 50 wt% PCL post-purification) and

DSC trace (Appendix F), elucidating a monomodal distribution and the existence of one

Tm, respectively.

Figure 3.2 – SEC trace for 50 wt% PCL shows free homopolymer. 50 wt% PCL

(post-purification) trace reveals homopolymer removal achieved with secondary

precipitation. Detection was based on differential refractive index.

Issues with the synthesis of these copolymers stems from the preference of ε-

caprolactone monomer to homopolymerize, rather than copolymerize from the sterically

hindered –OH moieties along the IIR backbone. Reaction conditions were varied (Table

3.1) in order to promote graft polymerization. However, the intended weight percentage

incorporation was not achievable due to inconsistencies with grafting and the non-

reproducible nature of the ROP.

57

3.1.2 Grafting of PCL onto the IIR Backbone

3.1.2.1 PCL Functionalization

Based on previous findings, in order to produce the target IIR-PCL graft copolymers, an

alternative, “grafting-to,” synthetic approach was required. Recent success in grafting

amine-terminated PEO to a IIR derivative having activated carbonates along the

backbone, indicated the same approach should be pursued. The first step was to prepare

an amine-terminated PCL. Using a previously reported method,217

an anhydride

derivative of di-tert-butyl dicarbonate (t-BOC)-protected β-alanine (3.6) was synthesized

(Scheme 3.2). As shown in Scheme 3.3, commercially available PCL-OH (900 g/mol,

3.7a, and 3500 g/mol, 3.7b) was then reacted with 3.6 to provide the protected amine-

functionalized PCL derivative 3.8a/b, which was deprotected with trifluoracetic acid

(TFA) to provide the target amine-functionalized PCL 3.9a/b. The synthesis of PCL-NH2

was confirmed by 1H NMR. The peak from the original PCL-OH polymer at δ = 3.65

ppm, representing the terminal methylene, shifted to 4.09 ppm upon reaction with 3.6 and

finally to 4.16 ppm after deprotection, as shown in Figure 3.3 (for 3.7a). Moreover,

methylene peaks corresponding to β-alanine appeared at 2.51 and 3.39 ppm, as well as a

methyl peak at 1.43 ppm belonging to the t-BOC group. The peak at 1.43 ppm

disappeared upon deprotection, and the methylene peaks of the terminal β-alanine

moieties shifted to 2.83 and 3.33 ppm, respectively. For functionalization of 3.7b, see

Appendix A.

Scheme 3.2 – Synthesis of t-BOC-protected β-alanine anhydride.

58

Scheme 3.3 – Functionalization of PCL (3.7a/b) by first reacting with BOC-

protected β-alanine (3.6) to produce the protected derivative (3.8a/b), followed by

deprotection with TFA (n = 8 for 900 g/mol PCL, 3.7a, initiated with ethylene glycol

derivative and n = 31 for 3500 g/mol PCL, 3.7b initiated with ethanol).

In addition, Tg and Tm were provided by Polymer SourceTM

(Table 3.2). This information

is important in order to observe how these temperatures are affected upon fabrication of

graft copolymers. SEC traces (Figure 3.4) showed that reaction of 3.7a and 3.7b with

BOC-protected β-alanine (3.6) and subsequent deprotection did not significantly impact

MWs, indicating that polymer backbone integrity was maintained. This was especially

important during TFA deprotection to ensure that ester hydrolysis did not occur.

Table 3.2 – PCL thermal properties (provided by Polymer SourceTM

).

Homopolymer MW (g/mol) Mn Mw PDI Tg (°C) Tm (°C)

3.7a 900 1872 2505 1.34 Not distinct 44

3.7b 3500 7621 9275 1.22 -64 63

59

Figure 3.3 – PCL (900 g/mol) functionalization (with k referring to the terminal

methylene): a) 3.7a; b) 3.8a; c) 3.9a.

Figure 3.4 – SEC traces of PCL derivatives throughout the functionalization

process: a) 3.7a – 3.9a (900 g/mol); b) 3.7b – 3.9b (3500 g/mol).

60

The synthesis of amine-functionalized PCL (3.9a/b) was also monitored by FT-IR in

order to observe N-H stretching in the 3300-3500cm-1

region. In congruence with

previously established work,218

typical stretching and vibrations associated with the PCL

homopolymer can be observed. Peaks at 2947 and 2866 cm-1

are due to CH2 vibrations,

the intense peak at 1728 cm-1

is due to C=O vibrations, CH2 bending vibrations at 1472,

1420 and 1366 cm-1

, C(O)-O vibrations at 1246 and 1171 cm-1

and lastly, C-O vibrations

at 1105 and 1047 cm-1

. However, after reacting 3.8a/b with BOC-β-alanine, N-H

stretching was observed at 3439 and 3393 cm-1

, as well as -NHCO- amide bond-

stretching at 1569 cm-1

. Moreover, after deprotection with TFA, N-H stretching appeared

as a single peak at 3445 cm-1

. It was important to ensure that peaks signifying the

presence of N-H stretching remained after deprotection, thereby confirming the

successful cleavage of the BOC group, liberating the amine-functionality for subsequent

copolymer fabrication (Appendix C for FT-IR traces).

3.1.2.2 Grafting of PCL to IIR

As shown in Scheme 3.4, IIR derivative 3.3 was activated with PNPC as per previous

methods (3.13).147

Following the Gillies group's protocol for the grafting of amine-

functionalized PEO, the amine-terminated PCL 3.9a/b was then reacted with the

activated IIR derivative (3.13) in the presence of DMAP at 60oC overnight (Scheme 3.5).

In order to purify each graft copolymer, redissolving in dichloromethane (DCM) with

subsequent water washing and multiple precipitations successfully removed residual

homopolymer (PCL) as well as impurities relating to 4-nitrophenyl carbonate. By

precipitating into acetone, it ensured that PCL polymers were removed. Yields were

typically in the range of 75-85%, with lower yields corresponding to graft copolymers

with higher percentages of aliphatic polyester incorporation. Increasing the fraction of

polyester incorporation could therefore increase the copolymers’ ability to dissolve in

acetone during purification.

Upon removal of ungrafted polyester homopolymers, the resulting graft

copolymers were characterized by 1H NMR, FTIR, DSC and SEC. Conversion of the

activated carbonates to carbamates upon successful grafting was revealed by shifts in 1H

NMR peaks (4.8-5.3 ppm) corresponding to the exo alkene and the C-H in the α-position

61

to the activated carbonate (Figure 3.5). In order to quantify PCL weight percentage, the

1H NMR peak corresponding to the PCL methylene triplet at 4.06 ppm (-CH2 adjacent to

oxygen) was integrated and compared against the PIB methylene (-CH2) peak at 1.41

ppm.

Scheme 3.4 – p-nitrophenyl chloroformate (PNPC) activated rubber synthesis.

Scheme 3.5 – Synthesis of PCL graft copolymers: 3.14 – 15 wt% PCL (n=8); 3.15 –

32 wt% PCL (n=31); 3.16 – 44 wt% PCL (n=31).

62

Figure 3.5 – 1H NMR spectra (CDCl3, 600MHz) of a) activated IIR; b) copolymer

3.15; and c) copolymer 3.16 showing how PCL content can be determined based on

the relative intensities of PCL to PIB, as well as reaction conversion determination

based on peaks from 4.8-5.3 ppm.

The appearance of a side-peak in SEC traces would indicate the presence of free

homopolymer. However, as observed in Figure 3.6, free PCL associated peaks did not

appear at their corresponding higher retention volumes (as compared to graft copolymers)

validating homopolymers were successfully grafted to the IIR backbone. However,

performing SEC analysis on IIR graft copolymers has been found to be problematic.132,148

Increasing PCL content should reflect an increase in MW, but this was not observed.

Instead, graft copolymers with increasing PCL content eluted at higher volumes;

polymers therefore behave anomalously on the column, as previously revealed in our

group by light scattering analysis. The SEC traces were limited as characterization tools

to ensure complete removal of ungrafted homopolymers, due to the inability to accurately

63

determine the MW data from these measurements. PCL content was controlled by

varying the MW of homopolymer (900 and 3500 g/mol) as well as the number of

equivalents relative to PNPC groups (1.0 and 0.8 equivalents of 3500 g/mol PCL-NH2) to

produce a small library of graft copolymers 3.14-3.16 (Table 3.3).

Figure 3.6 – SEC traces for: ungrafted IIR (3.1) and each IIR-PCL graft copolymer

(3.14-3.16).

Table 3.3 – IIR-PCL graft copolymers.

Copolymer PCL

MW

(g/mol)

Equivalent

(PCL -NH2)

Functionalized

Isoprene Units

(%)

PCL

Content

(wt%)

Mw

(kDa)

Tg

(°C) Tm

(°C)

3.14 900

1.2 100 15 504 -67 none

3.15 3500 0.8 85 32 395 -65 44

3.16 3500 1.2 100 44 458 -62 50

DSC measurements were also performed in order to determine Tg and Tm for each graft

copolymer. The Tgs of the graft copolymers were all very similar and were in the

expected range for both IIR (-70 °C) and PCL (-64 °C). Previous studies where PEO-IIR

graft copolymers were synthesized indicated that crystalline PEO homopolymer of 2000

g/mol, which displayed a Tm of 58°C, was significantly reduced upon incorporation into

the graft copolymer.148

Similar trends were found in the current study. Copolymer 3.16

(44wt% PCL), combining PCL homopolymer of 3500 g/mol with an initial Tm of 63 °C

was decreased to 50 °C upon covalent grafting to IIR. Copolymer 3.15 (32 wt% PCL),

with fewer equivalents of PCL (3500 g/mol) in relation to isoprene units exhibited a

64

further decrease in Tm to 44°C. Copolymer 3.14 (15wt% PCL), combining PCL

homopolymer of 900 g/mol with an intial Tm of 44°C, resulted in no apparent Tm for the

graft copolymer. Overall, the increase in melting temperature with increased PCL content

can be attributed to the ability of the larger PCL domains within these copolymers to

more readily crystallize in a manner that is similar to that of the homopolymer. DSC

traces would also show the presence of free, ungrafted PCL chains. In this case, there

would be an additional melting peak at the temperature for the corresponding PCL

homopolymer (Appendix H). This extra melting transition was not observed here,

confirming the purity of copolymers.

65

3.2 Grafting of PDLLA onto the IIR Backbone

3.2.1 PDLLA Functionalization

IIR-PDLLA graft copolymers were synthesized in a manner similar to that of the IIR-

PCL graft copolymers. First, as shown in Scheme 3.6, a commercially available

hydroxyl-terminated PDLLA 3.10 (2800 g/mol) was converted to the amine-terminated

PDLLA 3.12 using the same procedure described above for PCL. The only apparent

difference relates to the deprotection of PDLLA BOC-protected amine derivative 3.11

due to PDLLA’s apparent sensitivity to water. Successful deprotection of PDLLA to

afford the amine-terminated derivative (3.12) had to be performed using TFA under dry

conditions at a temperature of 0oC for 2 hours. The sensitivity to water could possibly

have caused the BOC protected β-alanine to be cleaved from the polymer terminus, and

additionally, a reduced temperature could have had a kinetic effect on the reaction,

thereby making it less favourable for the amine to be cleaved (when first attempted as per

PCL deprotection conditions). In this regard, upon deprotection, a cyclic lactam was

produced due to the cyclisation of the β-alanine amino acid. To reduce cyclisation, the

polymer was redissolved in DCM and passed over a K2CO3 plug in order to afford

PDLLA-NH2. Although conditions were slightly different (as compared to PCL-NH2),

the yields obtained were good and product purity was high.

Scheme 3.6 – Functionalization of 3.10 by first reacting with BOC-protected β-

alanine (3.6) to afford 3.11, followed by deprotection with TFA (3.12).

66

Figure 3.7 – Schematic depicting PDLLA functionalization: a) 3.10 (2800 g/mol),

starting material; b) 3.11, t-BOC protected β-alanine derivative; c) 3.12, amine-

liberated derivative. In 3.10-3.12, f represents the terminal methylene.

Successful synthesis of 3.12 was confirmed by 1H NMR spectroscopy. The terminal

methine peak found at δ = 4.36 ppm shifted to the 5.23 ppm region (overlapping with the

C-H α to the carbonyl) where it also remained after deprotection. Furthermore, methylene

peaks corresponding to β-alanine appeared at 2.53 and 3.42 ppm, as well as a methyl

peak at 1.43 ppm belonging to the BOC group (Figure 3.7). In addition, SEC was also

performed to ensure that there were no significant changes to the MW of the polymer

during this process (Figure 3.8). Finally, FTIR was performed on 3.10, 3.11 and 3.12 to

67

ensure the appearance of the peak corresponding to amine stretching in the 3500 cm-1

region, indicative of successful conversion of –OH termini to –NH2 termini of the

PDLLA polymer. Similar absorptions were observed in accordance with previously

reported PDLLA spectra.219,220

Vibration of the linear ester, carbonyl bands, C=O and

COO, appears at 1757 cm-1

, 1267 and 1134 cm-1

, respectively. In addition, characteristic

CH2 vibrations can be observed at 2997 and 2949 cm-1

. However, N-H stretching still

appeared upon reacting with BOC-β-alanine (similar to PCL functionalization) at 3517

and 3435 cm-1

(with amide –NHCO– bond-stretching at 1512 cm-1

) and remained after

deprotection as a single peak at 3508 cm-1

, thereby confirming successful cleavage and

liberation of amine-termini (Appendix C for spectra).

Figure 3.8 – SEC traces elucidating functionalization of PDLLA 3.10-3.12.

PDLLA starting polymer 3.10 was provided by Polymer SourceTM

and possesses a Tg of

28°C (Table 3.4). This information is important in order to observe how the glass

transition temperature is affected upon graft copolymer synthesis. As was expected,

PDLLA did not possess a melting temperature because of its structure; a racemic mixture

of D-lactide and L-lactide results in an amorphous polymer.

68

3.2.2 IIR and PDLLA Grafting

Using a procedure similar to that used for the synthesis of the PCL-IIR graft copolymers,

amine-terminated PDLLA 3.12 was reacted with the activated IIR derivative (3.13) to

prepare graft copolymer 3.17 (Scheme 3.7). The product was purified by multiple

precipitations into acetone. This not only successfully removed 4-nitrophenyl carbonate

impurities, but also removed residual PDLLA polymers. Yields in the range of 85-90%

were obtained. Graft copolymer 3.17 was characterized by 1H NMR, FTIR, DSC and

SEC. In congruence with copolymers 3.14, 3.15, and 3.16, conversion of the activated

carbonates to carbamates upon successful grafting was demonstrated by the shifts in 1H

NMR peaks (4.8-5.3 ppm) corresponding to the exo alkene and the C-H in the α-position

to the activated carbonate (Figure 3.9). PDLLA wt% was determined through integration

of 1H NMR corresponding to the PDLLA multiplet from 5.13-5.23 ppm (-CH α to

carbonyl), compared against the PIB methylene singlet at 1.43 ppm (-CH2). The PDLLA

content of copolymer 3.17 was found to be 30 wt% (Table 3.4). FT-IR spectroscopy

showed strong peaks associated with CH2 vibrations in the 3000 cm-1

region, arising from

both PIB and PDLLA, and an intense peak at 1700 cm-1

, characteristic of the PDLLA

C=O stretch (Appendix E). The purity of graft copolymer 3.17 was also demonstrated via

SEC analysis; Figure 3.10 displays a monomodal distribution, thereby confirming the

absence of free PDLLA homopolymer.

Scheme 3.7 – Synthesis of PDLLA Graft copolymer: 3.17 – 30 wt% PDLLA.

69

Figure 3.9 – PDLLA content in copolymer 3.17: determined via integration

corresponding to PDLLA multiplet from 5.13-5.23 ppm and PIB singlet at 1.41

ppm.

Table 3.4 – PDLLA homopolymer and IIR-PDLLA graft copolymer: PDLLA

content and thermal properties.

Polymer PDLLA

MW

(g/mol)

Equivalent

(PDLLA -

NH2)

Percentage

Functionalized

Isoprene Units

PDLLA

Content

(wt%)

Mw

(kDa)

Tg,1

(°C)

Tg,2

(°C)

Tm

(°C)

3.10 2800 4.7 28 none none

3.17 2800

1.2 100 46 462 -63 23 none

DSC analysis was also performed on copolymer 3.17. Contrasting with PCL, PDLLA,

because of its amorphous nature, does not exhibit a Tm (PLLA and PDLA are both

semicrystalline and do show Tms), rather just a Tg of approximately 28°C (Table 3.4). In

accordance with copolymer 3.14, the DSC trace of copolymer 3.17 exhibited a slight

change in Tg to -63°C from IIR’s Tg of -70°C. In addition, a Tg corresponding to the

PDLLA domains was observed at 23°C. This is a similar trend to what was observed with

IIR-PCL graft copolymers, however, the Tg of amorphous PDLLA (3.10) was affected.

Upon incorporation into the graft copolymer, the Tg of 3.10 was effectively reduced from

28°C to 23°C. Additionally, if present as a contaminant in the graft copolymers, the DSC

traces would also show the existence of free, ungrafted PDLLA chains. In this case, there

70

would be an additional glass transition peak at the temperature corresponding to 3.10’s Tg

(Appendix G). This extra glass transition was not observed here, confirming graft

copolymer purity.

Figure 3.10 – SEC traces of ungrafted IIR (polymer 3.1) and IIR-PDLLA graft

copolymer (3.17).

71

3.3 Preparation of IIR-PCL/PDLLA Blends

In order to observe the advantages that graft copolymers exhibit, physical blends

consisting of IIR and polyesters were also prepared. In brief, these blends were prepared

by dissolving IIR and either PCL or PDLLA in a common solvent, and then the solvent

was removed. To replicate the polyester content of the graft copolymers, blends with IIR

and 15, 32 and 44 wt% PCL as well as 30 wt% PDLLA were prepared and characterized

by 1H NMR and DSC. DSC traces (Appendix I) for blends consisting of PCL (5000

g/mol) and IIR were found to exhibit the same Tgs (-66°C) and Tms (50°C) regardless of

PCL wt% incorporation. It is also interesting to note DSC traces for PDLLA (18 000

g/mol) blends and IIR. PDLLA homopolymer was provided by Sigma-Aldrichwith a

glass transition temperature in the range of 38-42°C. In this case, two Tgs were observed,

one for IIR at -67°C and the other corresponding to PDLLA at 40°C (Appendix I).

72

3.4 Physical Characterization of Graft Copolymers

3.4.1 Atomic Force Microscopy

The study of IIR-polyester graft copolymer films by AFM was of interest in order to gain

insight into the phase separation and nanoscale morphologies of these polymers. In recent

work, Zhang et. al. have demonstrated that breakout crystallization occurs when studying

the thermal effects on block copolymers consisting of poly(butadiene)-block-PCL (PBD-

b-PCL).221

Breakout crystallization occurs when a crystalline portion that exhibits

nanometer length scale domains is subsequently heated, thus transforming it into

regularly alternating lamellae between the crystalline and amorphous layers. This

successfully alters (or destroys) the melt mesophase, due to the crystallization of one

block. Although the mechanism of breakout crystallization is poorly understood, it may

occur when the crystallization driving force is enough to overcome the energy barrier due

to the amorphous surroundings.222,223

Furthermore, when considering PBD-b-PCL, it has

been shown that PCL minority blocks do in fact break out into lamellar alternating

structures with PBD.224

Therefore, because IIR-PCL copolymers are similar to PBD-b-PCL (amorphous

and semicrystalline domains), AFM analysis could provide interesting images of

copolymer nanoscale structure pre- and post-annealing. Although it would be of interest

to understand the breakout kinetics (such as crystal coalescence and growth), AFM

equipment involving real-time imaging was not available. However, differences in

topographical and phase images could be analyzed, indicating that the development of

crystallization did in fact impact the surrounding amorphous microdomain structures. In

recent work, IIR-PEO graft copolymers exhibited micrometer scale patterns when spin-

cast on to silicon wafer surfaces.147,148

In this case, the patterning was attributed to

kinetic factors such as the freezing of Marangoni instabilities, as well as phase separation

(thermodynamically driven). In this sense, it would be interesting to study the pattern

formation with IIR-PCL graft copolymers. Although PCL, like PEO, is semicrystalline,

its inherent hydrophobicity could possibly affect the resulting surface morphology of

copolymers.

73

Thin films of copolymers 3.14-3.17 were prepared via spin-coating a 3 wt%

solution of copolymer in toluene onto silicon wafers. AFM showed that full surface

coverage was achieved; topographical and phase images were obtained and analyzed. As

shown in Figure 3.11, the surfaces were moderately rough at the nanometer scale. The

average roughness was found using XEI software to be 1.45, 0.394, 3.09 and 0.203 nm

for copolymers 3.14-3.17, respectively.

Figure 3.11 – Topography of copolymers: a) 3.14; b) 3.15; c) 3.16; d) 3.17.

IIR and PCL are both hydrophobic, but because they are chemically different,

phase separation is possible under certain conditions. Consistent with the lack of Tm in

DSC traces, no phase separation was observed for copolymer 3.14, containing 15 wt%

PCL (Figure 3.12a). However, upon increasing to 32 wt% PCL in copolymer 3.15, some

nanoscale patterning was observed (Figure 3.12b). By increasing the content of PCL

further to 44 wt% in copolymer 3.16, increased heterogeneity was observed (Figure

3.12c). From these results, it can be concluded that upon reaching a certain PCL content,

the PCL can separate from the melt and overcome the surface energy of IIR, which tends

to migrate to the top of the surface (lower surface energy). The repulsive energy from the

chemically different polymers must be great enough in copolymer 3.16 for phase

74

separation to occur, thereby producing significant microstructures and surface patterning.

In other words, free energy minimization during microphase separation therefore results

in interesting patterning.225

In the case of copolymer 3.17, the image was suggestive of

nanoscale phase separation, but well-defined patterns were not identified.

Figure 3.12 – Phase contrast of copolymers: a) 3.14; b) 3.15; c) 3.16; d) 3.17.

Upon annealing all of these surfaces, although partial organizing of PCL domains was

observed in copolymer 3.15, copolymer 3.16 displayed what appears to be a form of

breakout crystallization (Figure 3.13d). Regular and alternating domains of PCL and IIR

were formed on a nanometer scale. Perhaps the crystalline blocks were able to dissociate

out of their microdomains and into the crystal growth front. In addition, the spin-coating

process is kinetically driven, resulting in structures that are not thermodynamically

favoured. However, annealing allows the copolymers to form their most

thermodynamically favoured arrangement. Because the PCL chains are grafted at various

points in the IIR backbone (due to 0.5-4 mol%), phase separation is restricted to nano-

scale patterning. Therefore alternating lamellae of PCL form between IIR main chains.

Further studies would have to be performed in order to determine the nucleation,

75

coalescence and growth, however, as a preliminary study, terraced crystalline structures

did coalesce and form ordered structures upon crystallization of PCL domains.

Figure 3.13 – AFM analysis of copolymer 3.16. Before annealing: a) topography; b)

phase contrast. After annealing: c) topography; d) phase contrast.

Next, it was of interest to study the importance of the covalent grafting of PCL to the IIR

backbone by imaging the physical blends of IIR with PCL and PDLLA. Topographic

images proved that blended samples formed heterogeneous morphologies, characterized

by the formation of micrometer-scale polyester aggregates. The dark spherical portions

that can be observed in Figure 3.14, likely represent elastomeric particles dispersed

throughout both polyesters in prepared blends. In congruence with Gheno et. al,.226

acrylonitrile butadiene (NBR) with poly(vinyl chloride) (PVC) blends displayed spherical

aggregates that increased in size with increasing soft segment, NBR. From Figure 3.14a,

it is clear that the blend with the highest percentage of IIR (15 wt% PCL) displayed larger

spherical elastomeric aggregates (on average) when compared to polymer blends with

higher percentages of polyester. Moreover, in comparison to the graft copolymers,

polymer blends (with comparable polyester incorporation) showed an increase in surface

roughness to 5.78, 16.20, 4.78, 3.29nm for 15 wt% PCL, 32 wt% PCL, 44 wt% PCL and

76

30 wt% PDLLA blends, respectively. Likely, the covalent grafting of the polyester to the

rubber confines phase separation to the nanoscale, rather than micrometer scale.

Physically incorporated polyesters can phase separate at any length scale and are not

restricted by polymer dimensions. Knowing the composition of each blend also provided

information pertaining to regions within each image. It appears that the darker regions are

associated with the spherical elastomeric portions (depressions), forming by means of a

coalescence mechanism.227

Figure 3.14 – Topography images of IIR-polyester blends: a) 15 wt% PCL; b) 32

wt% PCL; c) 44 wt% PCL; d) 30 wt% PDLLA.

Understanding the morphology in these blends was important in order to properly

comment on the observed phase contrast images. In agreement with other rubbery-

semicrystalline blend systems, darker regions are typically associated with softer,

elastomeric portions.228

It should also be noted that PDLLA does differ from PCL due to

its amorphous nature. Because of this, PDLLA portions appear to represent darker

spherical regions, resulting due to the agglomeration of PDLLA. When observing the

phase contrast images in Figure 3.15a, b and d of 15 wt% PCL, 32 wt% PCL and 30 wt%

PDLLA, respectively, there appear to be regions outside of the major patterning, which

77

are likely attributed to blend portions consisting predominantly of IIR. However, Figure

3.15c, corresponding to 44 wt% PCL blend does not exhibit these regions. Having

reached a critical amount of PCL may have caused the crystalline portion to link itself,

thus forming tightly packed domains. This phenomenon has also been observed with

IIR/poly(butylene terephthalate) and liquid crystalline polymer blends in that increasing

the crystalline portion to 75% resulted in the formation of agglomerates and fibril

formation.227

Lastly, consistent with the graft copolymers, PCL blends of 32 wt% and 44

wt% do appear to show lamellar patterning of larger PCL domains. However, increasing

the weight percentage of the crystalline portion and allowing it more mobility caused an

increase of patterning and overall complexity of observed phase contrast images.

Figure 3.15 – Phase contrast of polymer blends: a) 15 wt% PCL; b) 32 wt% PCL; c)

44 wt% PCL; d) 30 wt% PDLLA.

78

3.4.2 Water Contact Angle Measurements

WCA evaluation is important when considering biomedical applications because

biomaterials will likely come into contact with water, blood, or other bodily fluids.

Understanding the material’s hydrophilicity and therefore wettability could impact its use

in certain applications. Table 3.5 illustrates the WCAs found for copolymers 3.14-3.17,

as well as PCL (900 g/mol, 3.7a and 3500 g/mol, 3.7b) and PDLLA (2800 g/mol, 3.10)

homopolymers. PCL-diol with a MW of 2000 g/mol has been shown to have a contact

angle of approximately 79°.229

Sessile drop analysis performed on PCL homopolymers

showed similar contact angles of 50.9° 1.77° and 70.8° 1.48°, for PCL of 900 g/mol

and 3500 g/mol, respectively (Table 3.5). However, increasing the MW of PCL shows

increased contact angles; PCL of 80 000 g/mol has displayed contact angles of 107°230

and 114°.218

Lower contact angles for lower MW polymers can likely be attributed to the

highly hydrophilic hydroxyl termini, which can have a larger impact in the context of

lower MW polymers. With PDLLA, similar trends were observed. Higher MW PDLLA

(125 000 g/mol) has been shown to have a static contact angle of 95.28° 0.18°.231

3.10

(PDLLA, 2800 g/mol) exhibited a lower contact angle of 66.1° 1.97° (Table 3.5).

Table 3.5 – Contact angle of PCL and PDLLA homopolymers and copolymers.

Homopolymer/Copolymer Contact Angle (deg)

3.7a 50.9

3.7b 70.8

3.10 66.1

3.14 92.1 0.68

3.15 91.5 2.30

3.16 94.1 1.38

3.17 91.0

79

Upon grafting 3500 g/mol PCL to the IIR backbone, copolymer 3.16 showed an increase

in contact angle to approximately 94°, whereas copolymer 3.15 showed an increase to

approximately 91° (relative to PCL homopolymer). Copolymers 3.14 and 3.17 had very

similar contact angles of 92 and 91° respectively. IIR has an approximate contact angle of

91°.232,233

Therefore the contact angles of the graft copolymers were very similar to those

of IIR. Although PCL and PDLLA homopolymers displayed lower contact angles, upon

functionalization and subsequent grafting, their hydroxyl termini were no longer

liberated. This would likely decrease their hydrophilicity, causing IIR to predominantly

influence the contact angle.

Contact angles of approximately 90° are within the range of typical contact angles

for various biomaterials.234,235

Implanted biomaterials can be problematic for thrombosis

formation and eliciting inflammatory responses, which are related to substrate wettability.

Although thrombus formation occurs through a biological system known as the

coagulation cascade, which is a healthy response of damaged tissues, thrombosis and

occlusion are particularly problematic when polymers are used to synthesize vascular

grafts.236

Thrombus formation begins with protein adsorption onto the foreign substance

and is more pronounced in terms of nonspecific protein binding onto highly hydrophobic

surfaces. Contact angles around 90° have not been shown to cause nonspecific protein

binding,237

therefore confirming their potential application for vascular prosthetics. In

order to confirm this, different techniques such as ellipsometry, quartz crystal

microbalance with dissipation monitoring and surface plasmon resonance can be

employed to determine protein adsorption.

80

3.4.3 Tensile Testing

Uncrosslinked IIR exhibits very low mechanical strength and low Young’s Modulus. It

was hoped that these properties could be improved through the incorporation of the

polyesters. Therefore, the mechanical properties of the new graft copolymers were

measured by tensile testing. The results of these tests are summarized in Table 3.6 and

the corresponding stress-strain curves can be found in Figure 3.16. PCL and PDLLA

have tensile strengths of 3.9 0.34 MPa (elongation at break (Eb) =61 and 41

MPa 6 MPa (Eb = 6.0-73.7% ), as well as Young’s Moduli of 83 9 MPa and

26.7 MPa 5 MPa, respectively.79,238-241

IIR has been shown to have E values of

approximately 0.20-0.5 MPa242

and a tensile strength of 0.09 MPa (Eb = 800%),243

and it

was of interest to determine how this was affected by grafting PCL and PDLLA to its

backbone.

By covalently combining both PCL and PDLLA with IIR, it provided rubber-

toughening, while also making the homopolymers more compliant by increasing their

resiliency. PCL and PDLLA are quite brittle, as can be observed by their relatively low

Ebs (i.e. brittle fracture). Graft copolymer synthesis allowed for a large increase in

elongation (Figure 3.16). As per Table 3.6, graft copolymer 3.14 experienced a 15-fold

increase in elongation whereas copolymer 3.17 showed a 42-fold increase (compared to

ungrafted PCL and PDLLA). Increases in elongation prove an increased compliance,

thereby decreasing brittle fracture that is experienced by the polyester homopolymers.

Additionally, graft copolymers experienced dramatic increases in Young’s Moduli

relative to IIR, which can be owed to an increase in crystallinity (due to PCL and

PDLLA). It was interesting to note that although copolymer 3.15 has a higher wt% of

PCL incorporation than copolymer 3.14, its stress vs. strain profile appeared to mimic

that of elastomeric IIR more closely (Figure 3.17 for a general elastomer profile).

81

Figure 3.16 – Stress-strain curve for: a) IIR; b) copolymer 3.14; c) copolymer 3.15;

d) copolymer 3.16; e) copolymer 3.17.

Table 3.6 – Tensile data of graft copolymers and IIR.

Copolymer Polyester %

in Copolymer

Young’s Modulus

E, at 50%

Elongation, MPa

Young’s

Modulus E,

MPa

Ultimate

Tensile

Strength

(UTS), MPa

Elongation

at Break

(Eb), %

IIR 0 0.66 0.19 0.56 0.12 0.247 ± 0.010 739 198

3.14 15 1.10 0.41 0.47 0.06 1.45 ± 0.20 1216 182

3.15 32 1.14 0.27 0.74 0.15 1.24 ± 0.13 447 65

3.16 44 9.57 1.13 21.66 5.90 3.92 ± 1.03 165 29

3.17 30 3.07 0.47 2.90 0.43 4.00 ± 0.86 251 52

By grafting PCL onto the IIR backbone, elongation was increased at varying

amounts depending on the weight percentage of PCL incorporation onto the IIR

backbone. The acquired data follows the expected trend; increasing the percentage of

PCL decreases its ability to elongate (relative to IIR), reaching a maximum decrease (7X)

by graft copolymer 3.16. Furthermore, E also increases from 0.47 MPa (copolymer 3.14)

to about 22 MPa (copolymer 3.16). Copolymer 3.16 exhibits somewhat brittle fracture (as

82

it behaves similarly to PCL homopolymer, Figure 3.17 for semicrystalline polymer),

whereas copolymers 3.14 and 3.15 display longer elongations and increased UTS

(compared to IIR). In addition, the tensile strength at break (Tb), or UTS, increased

dramatically from copolymer 3.14 and 3.15 to copolymer 3.16 by inducing a small

change in PCL content. Although it is difficult to conclude whether true TPEs have been

produced, confirmation of this could be accomplished via cyclic loading tests.

Figure 3.17 – Stress vs Strain of a variety of materials (with permission to reprint

from MIT OpenCourseWare, http://flic.kr/p/66XeQc).

Lastly, it is interesting to consider the study by Xu et. al. where they investigated

the influence of increasing the soft block in a PUR TPE.244

TPEs with higher hard

segment content (40-50% soft segment weight concentration, SSC) had higher moduli

and larger initial linear portion, whereas increased soft segment TPEs (60-70% SSC)

showed greater elastomeric behaviour and higher recoverability (Figure 3.18).

Mechanical properties of TPEs are generally described as materials exhibiting high

tensile strength and elongation of at least 2 times its original length.184

Although the

strengths of copolymers 3.14 and 3.15 were not significantly increased, they still display

excellent elongation properties with partial physical crosslinking. Their greater

elongation and better recoverability is owed to their more elastomeric behaviour.

Comparatively, copolymers 3.16 and 3.17 possess higher strengths with moderate

elongation. However, plasticization of the amorphous phase was still apparent. This is an

inherent characteristic of rubbery elastomers.245

83

Figure 3.18 – Thermoplastic PURs with varying soft segment contents. (Reprinted

from Polymer, 49/19, Xu et. al., Morphology and properties of thermoplastic

polyurethanes with dangling chains in ricinoleate-based soft segments (4248-4258).

Copyright (2008), with permission from Elsevier).244

From the AFM imaging, increasing to 44 wt% PCL in IIR copolymers induced

phase separation. Therefore, one can infer that the microphase-separated structure of the

IIR-PCL films enforced the mechanical characteristics, thereby imparting increased E

and UTS. The PCL domains that exist throughout the IIR matrix provide a physical

crosslinking mechanism, acting as fillers to strengthen the copolymer.243

Copolymer

3.17, when compared to IIR, also showed increases in Young’s Modulus and UTS from

0.56 – 2.90 MPa and 0.247 to 4.00 MPa, respectively. Comparatively, it does show

similar trends to PCL-based copolymers in that its overall strength is increased, however,

copolymer 3.17 (30 wt% PDLLA) has almost the same UTS as copolymer 3.16 (44 wt%

PCL). This is consistent with literature results in that PDLLA does display higher

strengths than PCL. Overall, homopolymers of PCL and PDLLA are extremely brittle.

For applications involving vascular grafts, compliance and a material’s ability to stretch

are of paramount importance.246

Therefore graft copolymer synthesis effectively tunes

and enhances the mechanical properties of IIR.

84

In biomedical applications, UTS and elongation are factors that govern suitability

of a material for a given application. For example, elastomeric materials display physical

properties that render them useful for implants involving soft or cardiovascular tissue.2,247

The vascular wall consists of elastin, collagen and smooth muscle with Young’s Moduli

ranging from 0.3-0.6 MPa, 100-2900 MPa and 0.006 MPa, respectively.107

Moreover,

elastin and collagen exhibit tensile strengths 0.36-4.44 MPa and 5-500 MPa. Therefore

for vascular stent applications, materials need to be extensible, allowing for vascular

dilation and constriction. Although mechanical properties of different soft tissue can be

mimicked by graft copolymers 3.14-3.17, it is important to note that many complex

factors dictate an effective vessel. For example, arteries are anisotropic, pulsatile,

compliant and thrombosis-resistant. Because of these complexities, in vivo testing of

possible prosthetics is critical when determining practical applicability.

Implantable materials need to reflect the mechanical behaviour (stresses/strains)

of the tissues that they are either replacing or supporting. It is apparent from Figure 3.19

that the diversity of toughness and E values of biological tissues requires a range of

biomaterials to be developed. Bone and tooth enamel require high moduli and moderate

UTS, due to their high-strength but brittle characteristics. However, collagen-rich tissue

such as intervertebral discs has moderate compliance and toughness because strength

needs to be maintained at higher strains. Figure 3.19 also depicts soft tissues with varying

levels of toughness and (lower) modulus combinations. However, successful clinical

implants typically have extremely high UTS (10-100MPa), which is much larger than

living tissue UTS (0.01-10MPa). Therefore, materials are needed with lower toughness

and moduli for cartilage, cardiovascular tissue and orthopaedic soft tissue. Copolymer

3.16, with a UTS of approximately 4 MPa and a Young’s Modulus of 22 MPa could

potentially serve as an implantable device for orthopaedic deficiencies, while copolymer

3.17 may be appropriate for softer cardiovascular tissue.

85

Figure 3.19 – Toughness-Modulus plot for current implant materials and the

mechanical regions for orthopedic hard tissue, orthopedic soft tissue, and

cardiovascular tissue. (Reprinted with permission from Taylor & Francis, 2013).246

Figure 3.20 – Stress-strain curve for: a) 15 wt% PCL blend; b) 32 wt% PCL blend;

c) 44 wt% PCL blend; d) 30 wt% PDLLA blend.

It was also important to demonstrate the superiority of covalent graft copolymer

mechanical properties when compared to simple, physical blends. As revealed by AFM

86

imaging, not having PCL or PDLLA covalently linked to the IIR backbone resulted in

free domains of polyester to be dispersed throughout the blend. Because domains were

randomly dispersed, thereby causing a lack in consistent ordering or patterning, this

manifested mechanical deficiencies, which were attributed to brittle PCL and PDLLA.

These deficiencies are predominantly characterized by the blends’ inability to withstand

higher stress loads (Table 3.7). Moreover, the blends were much weaker when compared

against their respective copolymer in terms of UTS. Figure 3.20 exhibits stress-strain

plots for PCL (15-44 wt%) and PDLLA (30 wt%) blends, showing what appears to be an

excess of noise, caused by the irregular composition of each blend. The blends’ yield

strengths are so high under tension that their yield strengths can never be reached because

the materials fracture first (linear regime followed by fracture, no plastic deformation). In

order to properly measure their yield strengths, unique tests that suppress fracture are

needed. One that may be employed is referred to as the compressive crushing strength;

however, this provides the elastic limit, σel, rather than the yield stress, σy.

Table 3.7 – Tensile data of polymer blends.

IIR –

Polyester

Blend

Polyester

% Relative

to IIR

Young’s

Modulus E, at

50% Elongation,

MPa

Young’s

Modulus E,

MPa

Ultimate

Tensile

Strength

(UTS), MPa

Elongation

at Break

(Eb), %

15 wt%

PCL

15 0.79 0.15 1.76 1.57 0.28 ± 0.038 516 114

32 wt%

PCL

32 0.95 0.18 2.29 2.06 0.18 ± 0.025 271 76

44 wt%

PCL

44 1.20 0.14 5.30 2.33 0.18 ± 0.043 141 52

30 wt%

PDLLA

30 N/A 15.11 4.51 0.62 ± 0.062 251 74

As the tensile testing was performed on each blend, the sample seemed to fracture at the

polyester domains; sample remained was still pulled in the axial direction due to IIR fibre

attachment. The difference between many other studied IIR and thermoplastic polymer

blends is typically the IIR is vulcanized prior to/during the blending process.248

In these

cases, blends showed an increase in UTS and E.

87

It can be concluded that the graft copolymers were superior in terms of their

strength (compared to IIR and blend systems), which is an important property when

considering materials for implantable biomedical devices. Although graft copolymers and

blends were non-crosslinked (unvulcanized), the strength and rigidity added by the

aliphatic polyester covalent attachment may serve as an appropriate avenue for avoiding

harsh vulcanization conditions. This has been coined as ‘green strength’: the strength,

cohesiveness, dimensional stability, and extensibility of rubber compounds prior to

curing/vulcanization.228

By avoiding vulcanization, graft copolymers may be considered

viable candidates for in vivo applications.

88

3.5 Degradation Study of IIR-PCL/PLA Graft Copolymers

PCL and PDLLA have both been shown to degrade particularly slowly. Under

physiological conditions, PCL has been found to degrade over a period of 3-4 years,84

and PDLLA degrades over approximately 12-16 months.73,76

IIR possesses high chemical

stability, to the extent where it shows little to no degradation over several years. Because

PCL and PDLLA were grafted onto the IIR backbone, it was expected that their

degradation rates would be slowed. Since degradation rates were expected to proceed

exceedingly slow, the focus of this study was therefore directed toward analyzing

accelerated degradation rates by subjecting samples to 5M NaOH over a four month

period.249-251

3.5.1 Mass Evolution and Scanning Electron Microscopy

Interestingly, the accelerated degradation of copolymers 3.14-3.16 resulted in

minimal weight loss, on the order of 0.5-1% change in weight (Figure 3.21, Appendix J).

However, such minimal change in weight could be due to experimental discrepancies

resulting from weight measurement, associated with the detection limit and low initial

masses. In order to correct for this, future experiments should employ samples of larger

mass to realize greater differences pertaining to crude mass loss. However, these results

were still quite significant when comparing to those of the control materials. Both IIR as

well as PCL were studied under the same accelerated conditions to observe what affect

the copolymerization of these materials had on their apparent degradation characteristics.

The PCL control (MW approximately 5000 g/mol) fully degraded in less than 24 hours.

In previous work, it has been found that the longer (and therefore higher MW) the

aliphatic polyester was, slower degradation rates would therefore result.142,252

Although

mass loss was not apparent, the shorter PCL grafted chains of 900 g/mol (copolymer

3.14) in comparison to 3500 g/mol (copolymers 3.15 and 3.16) did show an increased

rate of hydrolytic degradation, apparent in the subsequent SEC analysis section.

89

Figure 3.21 – Mass loss of copolymers 3.14-3.16, PCL and IIR controls.

The accelerated degradation of copolymer 3.17 was also studied to gauge if

grafting of PDLLA to IIR exhibits accelerated degradation rates when compared to IIR-

PCL copolymers. Due to time limitations, this study has only proceeded to the 2-month

time point. Crude mass loss results were similar to copolymers 3.14-3.16 in that minimal

weight loss, approximately 0.5-1.0 wt%, was observed (Figure 3.22, Appendix K for raw

data). The same reasoning suggests that low perceived weight loss could be due to

experimental discrepancies associated with the detection limit and low initial masses of

testing materials. In congruence with PCL homopolymers, PDLLA showed complete

degradation. However, full degradation was achieved in 5 days (compared to 24 hours),

attributed to its higher MW (18 000 g/mol) than PCL (5000 g/mol).

90

Figure 3.22 – Mass loss of copolymer 3.17 as well as PDLLA and IIR controls.

Next, when studying SEM results, PCL concentration appeared to significantly

impact graft copolymer characteristics. As PCL wt% was increased, the obtained surfaces

exhibited an increase in plasticity. It can be inferred that graft copolymers manifested

partial TPE properties due to their excellent processability. When the discs were prepared

via melt-pressing at elevated temperatures, the hard phase of the copolymers (PCL)

would melt. However, upon cooling, the hard phase would solidify, producing

copolymers with strength and elasticity. Copolymers exhibiting a greater portion of the

hard phase therefore solidified into samples with reduced flowability, i.e. less elastomeric

portion. Therefore copolymer 3.16 possessed higher plasticity resulting in flat and

smooth topographies, which remained highly unchanged throughout the duration of the

study (Figure 3.23d) and h)). Although copolymer 3.17 consists of PDLLA (not a

semicrystalline “hard” phase), it does possess high UTS and E that are comparable to

PCL. Perhaps these properties also caused copolymer 3.17 to have and retain smooth

topographies over a 2 month period (Figure 3.24a and b). After just one month, IIR and

copolymers 3.14 and 3.15 showed wrinkling, similar to their 4 month time points (Figure

3.23). Surface irregularities, however, were particularly pronounced for copolymers 3.14

and 3.15, resulted in corrugated surfaces. This is perhaps attributed to the degradation of

PCL. These findings also coincide with the notion of higher MW aliphatic polyesters

displaying a decreased degradation rate. A higher degree of PCL incorporation and MW

91

for copolymer 3.16 (compared to copolymer 3.14) suggests that there is a relationship

between polyester concentration and the ability to maintain surface integrity.

Figure 3.23 – SEM imaging taken at 100X magnification under variable pressure

mode at To and 4 months of: a) and b) IIR; c) and d) copolymer 3.14; e) and f)

copolymer 3.15; g) and h) copolymer 3.16.

92

Copolymers 3.14 and 3.15 and IIR displayed creep deformation resulting in

sample wrinkling. Macroscopic images qualitatively portrayed the aforementioned

sample deformation (Figure 3.25). As it can be seen, although there is no apparent weight

loss, samples with higher amounts of IIR tend to shrink as they are subjected to an

aqueous environment. Factors governing creep wrinkling can be attributed to differences

in mechanical properties of IIR and PCL, resulting in smooth, shallow surface

undulations due to uneven material expansion.253

Moreover, wrinkling has been shown to

relax the compressive strain in the hard layer (PCL), thereby reducing elastic strain

energy.254

Finally, the four month time point for copolymers 3.14 and 3.15 resulted in

opaque materials, where the starting materials were more translucent. As PCL degraded,

the crystallinity of the resulting degraded chains may have caused a decrease in sample

transparency.

Figure 3.24 – SEM imaging taken at 100X magnification under variable pressure

mode at To and 2 months of: a) and b) copolymer 3.17. Macroscopic images of To-2

months of: c) copolymer 3.17.

93

The macroscopic images are particularly interesting as there as some notable

characteristics pertaining to copolymer 3.17 (Figure 3.24c) and 3.16 (Figure 3.25d).

Wrinkling was almost negligible, along with maintenance of is translucent nature. These

copolymers were not affected by creep deformation, which is favourable for different

implantable applications such as vascular prosthetics. High amounts of degradation for

such applications would not be favourable as they require mechanical strength and

rigidity, along with structural integrity. Although copolymers 3.16 and 3.17 do not have

as much strength or toughness as PIB-PMMA block copolymers, it is this decrease in

rigidity yet structural integrity that makes them more suitable for soft-tissue-based

applications (wound healing, sutures, intervertebral disc and articular cartilage repair).246

Figure 3.25 – Successive macroscopic images representing To, 1 month, 2 months, 3

months and 4 months of: a) IIR control; b) copolymer 3.14; c) copolymer 3.15; d)

copolymer 3.16.

There is a large interest in TE applications therefore synthesizing materials

incorporated with biodegradable chemistries and materials that can withstand strains of

up to 100% is highly favourable.255,256

It is the combination of the ability to maintain

mechanical integrity, elongation strains and appropriate strength that makes a material

suitable as a shape memory polymer (SMP). Although SMPs can be synthesized across a

wide range of polymer chemistries including poly(methacrylates/acrylates), aliphatic

94

polyesters and PURs, there are several degradation and mechanical related issues.257-259

Original SMPs were based upon polymers such as PLA, PCL and poly(glycolide), with

subsequent methacrylation to allow for free-radical polymerization, procuring crosslinked

networks for drug delivery applications.260

These SMPs suggested that although initial

degradation did not affect material properties, degradation at amorphous regions

increased crystallinity (and therefore E), thereby increasing brittle behaviour. However,

degradation at the crystalline domains softened and decreased the modulus of the

materials, resulting in structural deficiencies. Therefore, copolymers 3.16 and 3.17’s

ability to maintain their structural integrity with average elongations at break of 165%

and 251% respectively (Table 3.6), demonstrates their applicability for applications

requiring structural maintenance and resiliency. A decrease in rigidity and maintenance

of structural integrity is the greatest accolades a polymeric device for such specific

biomedical applications can therefore possess. In vivo conditions would have to be

further examined as this may compromise structural integrity (cyclic mechanical loading

and biochemical attack). Although these resulting deficiencies are more critical for

polymers containing hydrophilic sensitivities, hydrophobic polymers may still experience

a loss in toughness due to internal crazing effects or gradual surface hydrolysis.261

However, PCL and PDLLA reinforced with IIR provide adequate water penetration

resistance.

95

3.5.2 Size Exclusion Chromatography

In addition to crude mass loss measurements, SEC was also performed on polymeric

materials at each time point. Even though there were no significant changes in mass,

hydrolytic degradation under accelerated basic conditions may have affected copolymer

MW profiles. Observing changes in MW will provide further insight toward

understanding copolymer properties and applicability as biomaterials. The graphical

representations shown in Figure 3.26 depict the change in Mn and Mw over the 16-week

study period for copolymers 3.14-3.16 and 8 week period for copolymer 3.17. Figure

3.26a shows that the Mn of IIR (control), copolymer 3.16 and copolymer 3.17 remained

highly unchanged, which is attributed to IIR’s chemical stability and copolymer 3.16 and

3.17’s increased polyester concentration.

Figure 3.26 – MW data for IIR control as well as copolymer 3.14, 3.15, 3.16 and

3.17: a) change in Mn; b) change in Mw.

96

Copolymers 3.14 and 3.15 showed gradual decreases in Mn values (Figure 3.26a). In

congruence with this finding, Figure 3.27b shows a significant shift in the MW profile of

copolymer 3.14. Lower MW polyester chains will degrade faster; copolymer 3.14,

prepared with the lowest MW polymer (900 g/mol PCL), therefore exhibited the largest

decrease in MW. Since it was prepared with lower MW PCL, cleavage of PCL chains

underwent further degradation, resulting in oligomers that possibly diffused out of the

polymer matrix (absence of lower MW side-peak in SEC traces). A simultaneous

decrease in Mw was also realized, causing a dramatic but controlled decrease in MW

profile.

Figure 3.27 – MW profiles elucidating each time point over the 4 month study

period for: a) IIR control; b) copolymer 3.14; c) copolymer 3.15; d) copolymer 3.16.

Even though copolymer 3.15 did show a decrease in Mn, its Mw did not show a

corresponding decrease, causing the main peak (attributed to IIR) to remain stationary

97

(Figure 3.27c) causing a significant increase in PDI. This behaviour is attributed to the

shedding of the lower MW PCL side chains and the formation of a bimodal polymer

distribution.

Figure 3.28 – MW profile elucidating each time point over the 2 month study period

for copolymer 3.17.

Copolymer 3.16 did not show a significant decrease in either Mw or Mn, which is

attributed to higher PCL concentrations. Miao et. al. showed that increasing the level of

PCL and therefore crystalline domains in triblock copolymers of PCL and poly(sebacic

anhydride) resulted in slower degradation profiles in both physiological and basic

conditions.262

Therefore in this thesis, increasing PCL content relative to IIR may be the

determining factor for decreased degradation. Conversely, increasing amorphous domains

allows water to penetrate and degrade materials easier. Because of this, degradation of

IIR-PDLLA copolymer 3.17 was hypothesized to display an increased weight loss and

change in MW due to PDLLA’s amorphous state. However, Figure 3.28 showed that its

MW profile remained relatively constant over the first 8 weeks of the study.

Degradation of PCL can occur via surface or bulk processes. However, a

combination of these pathways may have influenced IIR-PCL copolymer degradation.90

Prediction of erosion pathways is particularly complicated when grafting polyesters that

typically abide by bulk erosion (where degradation is unpredictable and starts

spontaneously) onto IIR (one that does not show degradation).263

Although mass

remained almost constant for all samples, Mn did decrease for copolymers 3.14 and 3.15,

98

indicative of PCL chain cleavage and subsequent oligomer formation.264,265

Chain

scission was homogeneous for copolymer 3.14 causing a homogenous decrease in MW

(bulk degradation). It is interesting to note, however, that 3.16 (44 wt% PCL

incorporation) showed little to no change in MW. This is similar to surface degradation in

that water cannot penetrate the polymer, causing the hydrolysed byproducts to diffuse

rapidly into the media, resulting in limited water penetration to the copolymer matrix. As

a result, degradation will occur purely at the surface, causing the polymer to thin while

leaving the MW intact. However, under base-accelerated conditions, autocatalysis at the

surface would not occur. Therefore the high concentration of PCL combined with stable

IIR caused water penetration to be negligible resulting in minimal degradation.

99

3.6 Bioassays and Compatibility

3.6.1 Cell Growth on Polymer Films

The evaluation of cell growth on copolymer surfaces was performed by studying the

adhesion and topology of C2C12 murine myoblast cells on films of copolymers 3.14-

3.17, and comparing them to the same cells grown on selected control surfaces. Although

this method is qualitative, observing cell morphology is an essential screening method for

cytotoxicity and cell compatibility at the initial analysis stage.191

Observation of how the

cells behave in terms of adhesion and viability gauges cellular response, which can be

confirmed via the MTT (3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyltetrazolium bromide)

cytoxicity test.191,266

To prepare the films for these studies, copolymers 3.14-3.17 were

prepared as 3 wt% solutions in toluene with subsequent drop-casting onto glass

coverslips, followed by sterilization.

Seeding myoblasts onto polymer surfaces and comparing cell proliferation to

glass, IIR, PCL and PDLLA homopolymers, allowed for direct comparisons between

substrate and cell morphology. After allowing an incubation period of 48 hours, cells

were stained with DAPI (4’-6-diamidino-2-phenylindole dyhydrocholoride) and Alexa

Fluor 568 phalloidin in order to visualize nuclei and actin protein fibers of the

cytoskeleton, respectively. Observation of the cytoskeleton structure is crucial.

Scrutinizing cellular spread and resulting morphologies provides qualitative comparisons

between copolymers and control substrates to assess cellular health.267

In addition, cells

were also counted on each surface for a quantitative measure. In order to do so, surfaces

were prepared in triplicate for each sample and 3 images were taken at random locations

to determine cell density. Utilizing cellc12 and Matlab 9 numerical values were obtained.

By averaging these values, determining standard deviations, and performing ANOVA

statistical measurements (Prism), confirms whether cell densities are/are not significantly

different between each surface. Figure 3.29 shows representative images from the control

samples (a-d) as well as from each copolymer (e-h). In this regard, one can clearly

evaluate the results on a qualitative and quantitative basis.

100

Figure 3.29 – Growth of murine myoblast cells on: a) glass; b) IIR; c) PCL; d)

PDLLA Cell imaging of copolymers: e) 3.14; f) 3.15; g) 3.16; h) 3.17.

Qualitatively, the spread of the cytoskeletons for graft copolymers in Figure 3.29

are comparable to those of the controls. A healthy cytoskeleton effectively spreads across

the surface, ensuring proper adhesion of the cell to the test material. However, on the

glass coverslip, as well as PDLLA and PCL controls, cells generally appeared smaller

and less well-spread. Differences in morphology may be owed to differing degrees of

surface roughness; smoother surfaces (such as the controls mentioned) lead to differing

focal adhesions. It has been previously reported that uneven biomaterial surfaces (with

higher surface roughness) can cause cells to grow differently, thereby affecting their

surface response.202

In addition, recalling that the surfaces of copolymers are somewhat

hydrophobic (about 90°), this may also impact the differences in cell adherence when

comparing between PDLLA, glass coverslip and PCL controls (much more hydrophilic).

More hydrophobic surfaces seemed to result in cells that were more flat, bipolar or

tripolar in morphology, and also formed lamellipodia (actin projection on mobile edge),

which is suggestive of high mobility and strong adhesion.266

This may be in part

attributable to the rapid adsorption of proteins from the cell culture media onto the more

hydrophobic surfaces, providing an ideal surface for cell adhesion. Stress fibers were also

imaged on all surfaces, implying that cell growth and proliferation was successfully

demonstrated and that actin cytoskeleton organization occurred. Actin stress fibres are

101

typically associated with strong focal adhesions, which could be tested via vinculin

labeling (major protein associated with focal adhesions). However, although cells were

found to be more rounded on surfaces with slightly higher hydrophilicity, increasing the

incubation time did result in increased cell elongation and therefore adhesion. Since

C2C12 murine myoblast cells proliferate quickly (doubling in approximately 27 hours),

some confocal images (such as Figure 3.29e) featured cells that displayed tips deeply

stained by the dye, indicative of dynamic contact. Overall, actin filaments were observed

on all tested surfaces. This linear polymer microfilament is important for cellular

functions such as mobility and contraction of cells during division, thereby confirming

the existence of healthy, proliferating cells.

Quantitatively, unhealthy cell specimens would result in much fewer cells

adhering to the substrate if conditions were unfavourable for proper attachment. Figure

3.30 reveals the average density of cells on each surface, which ranged from 270-400

cells/mm2. ANOVA tests confirmed that there were no significant differences between

the results for the different surfaces. Control substrates such as glass are considered to be

reasonably good substrates for cell growth, indicating that the graft copolymers provide

favourable conditions as well. These results are crucial in determining cell adhesivity

properties for potential implantable biomaterial applications. This importance is related to

surgical implantation of different medical devices in that local tissue injury resulting from

insertion of vascular grafts, heart valve sewing cuffs and annuloplasty rings, affects both

thrombosis and inflammation at the site. Both of these processes are also known to

contribute to healing of tissue into and around the device.268

Ensuring that a monolayer of

endothelial cells can form on the surface of vascular grafts, for instance, can prevent

thrombosis and allow for faster healing times.108

Moreover, this preliminary test also

suggests that copolymers will provide a non-toxic environment, which will be confirmed

via the MTT cytotoxicity assay.

102

Figure 3.30 – Cell adhesivity quantified by determining average cell concentrations

for each substrate.

103

3.6.2 MTT Toxicity Assay

Performing cytoxicity testing is also considered a pre-screening test; it is essential

to carry out in order to quantify the number of cells that will remain viable upon

incubation with potentially toxic biomaterial leachables. Fatal leachables may diffuse out

of a material, thus causing a toxic environment, which is detrimental to cell proliferation.

This test allowed for quantification of cell growth and differentiation associated with

controls [glass, IIR, PCL, PDLLA and high density poly(ethylene)] and copolymers 3.14-

3.17. A slight adaptation of ISO 10993-5, ‘Biological Testing of Medical Devices – Part

5: Tests for Cytoxicity, in vitro method, tests on extracts,’ was used as a quantitative

means for determining cell viability. The ‘extract’ method was used in order to determine

toxic effects that could possibly be generated by the biomaterials. In congruence with the

American Society for Testing and Materials, approximately 80 mg of the material was

used for each test. Melt-pressing the material into a uniform surface and cutting into 1

cm2 pieces maintained surface area exposure for reproducible and reliable results. The

culture media was used as an extracting medium at 37 °C for 24 hours prior to testing, as

this allows sufficient time for possible leachables to enter the surrounding media. The

cells were then incubated in serial two-fold dilutions of the leachate with fresh culture

media. After 24 h incubation time with cells, a standard MTT cell viability assay was

performed. In this case, a reduction of cell viability by more than 30% relative to control

cells exposed only to fresh culture media is considered to be a toxic effect.

The results of the MTT assay are shown in Figure 3.31. Positive and negative

controls were included in each of the tests. High density poly(ethylene) (HDPE) was used

as the negative control, as it does not elicit cytotoxic effects. Sodium lauryl sulfate (SDS)

was used as a positive control to cause cell lysis, thereby compromising its integrity and

health. Although PCL, PDLLA and IIR polymers are known to be biocompatible,

investigating copolymer cytotoxicity and comparing between controls is highly relevant.

As indicated in Figure 3.31, neither the negative control (HDPE) nor graft copolymers

3.14-3.17 exhibit any significant cytotoxic effects. However, toxicity of SDS was

confirmed at the expected concentrations of 0.2-0.10 mg/mL.

104

Figure 3.31 – MTT cytotoxicity assay performed on: a) graft copolymers 3.14-3.16;

b) graft copolymer 3.17.

105

Chapter 4

4 Materials and Methods

4.1 General Procedures and Materials

IIR 402 with a Mw of 395kDa and a polydispersity index (PDI) of 2.44, as measured by

size exclusion chromatography (SEC), composed of 2.2mol% isoprene units was kindly

provided by LANXESS and was converted to the activated derivative (3.13) by the

previously reported method.147

Hydroxy-terminated PCL (MW = 900 Da and 3500 Da)

and PDLLA (MW = 2800Da) were purchased from Polymer Source (Montreal, Québec).

Silicon wafers were purchased from University Wafer (Boston, Massachusetts). Cell

culture materials were purchased from Nunclon and Invitrogen. Solvents were

purchased from Caledon, PNPC was purchased from Alfa Aesar (used as received). m-

chloroperbenzoic acid was dissolved in toluene and dried with MgSO4 before use.

Pyridine and ε-caprolactone monomer were distilled over CaH2 before use. All other

chemicals were purchased from Sigma Aldrich and used without further purification

unless stated otherwise. Dry dichloromethane and toluene were obtained from an

Innovative Technology (Newburyport, USA) solvent purification system based on

aluminum oxide columns. 1H NMR spectra were obtained in CDCl3 at 600 MHz using

Varian Inova spectrometers. NMR chemical shifts are reported in ppm and are calibrated

against residual solvent signals of CDCl3 (δ 7.26). Coupling constants (J) are reported in

Hz. The percentage of functionalized isoprene units was determined from 1H NMR,

based on the relative integrations of the signals at 5.03 and 4.87 ppm coinciding to the

alkene adjacent to the activated carbonate and the PCL/PDLLA carbamate product,

respectively. The PCL and PDLLA content in weight percentage was determined by 1H

NMR, based on the relative integrations between the peaks at 4.03ppm (PCL methylene),

5.19ppm (PDLLA methane) and 1.12ppm (isobutylene). Differential

scanning calorimetry (DSC) was performed under a nitrogen atmosphere on a Q20 DSC

TA instrument at a heating/cooling rate of 10 °C/min from -100 to + 100 °C. The Tg and

Tm were obtained from the second heating cycle. SEC was performed in tetrahydrofuran

(THF) using a Viscotek GPCmax VE 2001 GPC Solvent/Sample Module equipped with a

106

Waters 2489 UV/Visible Detector, Viscotek VE 3580 RI Detector and two PolyPore (300

mm x 7.5 mm) columns from Agilent. The calibration was performed using polystyrene

standards. FTIR was performed on a Bruker Optics TENSOR 27 series FT-IR, OPUS 7.0,

via transmittance (%) with a background scan of 16 and a sample scan of 128, recording

wavenumbers from 500-3700cm-1

.

4.2 Graft Copolymer Synthesis and Chemical Characterization

4.2.1 Synthesis of Polymer 3.8a

Following a previously reported literature,148

a round-bottom flask equipped with stir bar

was charged with PCL homopolymer of 900 g/mol (3.7a) (0.40 g, 0.44 mmol, 1.0 equiv.),

4-(dimethylamino)pyridine (DMAP) (0.20 g, 1.64 mmol, 3.7 equiv.) pyridine (0.10 g,

1.29 mmol, 2.9 equiv.) and dichloromethane (DCM) (7 mL) under dry conditions. BOC-

protected β-alanine anhydride (3.6) (0.39 g, 1.1 mmol, 2.5 equiv.) was charged into a

separate flask and dissolved in DCM (2 mL) under dry conditions. The latter solution was

added to the PCL-containing solution and stirred overnight at room temperature. Next,

deionized water was added to the reaction mixture and stirred for an additional 3 hours at

room temperature. Purification involved washing with 1M HCl (3X), 1M Na2CO3 (3X)

and concentrated brine (1X), followed by drying with MgSO4 and reduced pressure

solvent removal.

Yield: 0.35 g, 83% 1H NMR: δ, 4.24 (t, 2H, J = 4.7 Hz), 4.04-4.10 (m, 18H), 3.70

(t, 2H, J = 5.1 Hz), 3.63-3.66 (m, 2H), 3.54-3.56 (m, 2H), 3.40 (br. s, 2H) 3.39 (s, 3H),

2.51 (t, 2H, J = 6.1 Hz), 2.29-2.37 (m, 16H), 1.60-1.69 (m, 32H), 1.43 (s, 9H), 1.34-1.42

(m, 16H). SEC: Mw = 3160 g/mol, PDI = 1.33. IR: 1047, 1105, 1171, 1246, 1366, 1420,

1472, 1569, 1728, 2947, 2866, 3393, 3439 cm-1

.

Polymer 3.8b was prepared using the same procedure as described for polymer 3.8a,

expect that PCL homopolymer of 3500 g/mol (3.7b) (0.40 g, 0.114 mmol, 1.0 equiv.) was

used.

107

Yield: 0.38 g, 88% 1H NMR: δ, 4.13 (td, 2H, J = 7.8, 4.7 Hz), 4.06 (t, 62H, J = 6.8

Hz), 3.39 (q, 2H, J = 6.1 Hz), 2.51 (t, 2H, J = 6.1 Hz), 2.3 (t, 62H, J = 7.4 Hz), 1.62-1.68

(m, 124H), 1.43 (s, 9H), 1.34-1.42 (m, 62H), 1.25 (t, 3H, J = 7 Hz). SEC: Mw = 9202

g/mol, PDI = 1.22. IR: 1047, 1105, 1171, 1246, 1366, 1420, 1472, 1569, 1728, 2947,

2866, 3393, 3439 cm-1

.

Polymer 3.11 was prepared using the same procedure as described for polymer 3.8a,

expect that PDLLA homopolymer of 2800 g/mol (3.10) (0.40 g, 0.143 mmol, 1.0 equiv.)

was used.

Yield: 0.31 g, 81% 1H NMR: δ, 5.12-5.25 (m, 39H), 4.23-4.32 (m, 2H), 3.57 (m,

2H), 3.42 (br. s, 2H), 3.36 (s, 3H), 2.59 (m, 2H), 1.54-1.59 (m, 117H), 1.43 (s, 9H). SEC:

Mw = 5098 g/mol, PDI = 1.19. IR: 1134, 1267, 1512, 1757, 2949, 2997, 3517, 3435 cm-1

.

4.2.2 Synthesis of Polymer 3.9a

Polymer 3.8a (0.34 g, 0.373 mmol, 1.0 equiv.) was dissolved in 1.25 mL DCM: 1.25 mL

trifluoroacetic acid (TFA) (1:1) and the reaction mixture was stirred for 2 hours. DCM

and TFA were removed by pressurized air. The product was redissolved in DCM and

dried in vacuo.

Yield: 0.34 g, > 99% 1H NMR: δ, 4.24 (t, 2H, J = 4.7 Hz), 4.16 (t, 2H, J = 6.4

Hz), 4.06 (t, 16H, J = 6.6 Hz), 3.7 (t, 2H, J = 4.7 Hz), 3.64-3.66 (m, 2H), 3.55-3.58 (m,

2H), 3.39 (s, 3H), 3.33 (br. s, 2H), 2.8 (t, 2H, J = 5.9 Hz), 2.29-2.34 (m, 16H), 1.60-1.69

(m, 32H), 1.34-1.42 (m, 16H). SEC: Mw = 2450 g/mol, PDI = 1.45. IR: 1047, 1105, 1171,

1246, 1366, 1420, 1472, 1569, 1728, 2947, 2866, 3445 cm-1

.

Polymer 3.9b was prepared using the same procedure as described for polymer 3.9a,

expect that polymer 3.8b (0.40 g, 0.114 mmol, 1.0 equiv.) was used.

Yield: 0.40 g, > 99% 1H NMR: δ, 4.17 (t, 2H, J = 6.5 Hz), 4.10-4.15 (m, 2H),

4.06 (t, 62H, J = 6.5 Hz), 3.33 (t, 2H, J = 5.9 Hz), 2.83 (t, 2H, J = 5.3 Hz), 2.31 (t, 62H, J

= 7.6 Hz), 1.62-1.68 (m, 124H), 1.36-1.41 (m, 62H), 1.25 (t, 3H, J = 7.6 Hz). SEC: Mw =

108

9030 g/mol, PDI = 1.22. IR: 1047, 1105, 1171, 1246, 1366, 1420, 1472, 1569, 1728,

2947, 2866, 3445 cm-1

.

4.2.3 Synthesis of Polymer 3.12

Polymer 3.11 and 2 mL of DCM were charged in a round-bottom flask equipped with stir

bar under dry conditions. TFA was added (0.70 g, 6.11 mmol, 100 equiv.) and stirred at

0°C for 2 hours. Next, DCM and TFA were removed under low pressure (~20 mbar), re-

dissolved in DCM and passed over a K2CO3 plug to remove residual TFA.

Yield: 0.65 g, 91% 1H NMR: δ, 5.13-5.23 (m, 39H), 4.23-4.32 (m, 2H), 3.57 (tt,

2H, J = 2.35, 3.52Hz), 3.36 (s, 3H), 3.29 (br. s, 2H), 2.83 (br. s, 2H), 1.54-1.58 (m,

117H). SEC: Mw = 4740 g/mol, PDI = 1.31. IR: 1134, 1267, 1512, 1757, 2949, 2997,

3508 cm-1

.

4.2.4 Synthesis of Graft Copolymer 3.14

In a round-bottom flask equipped with stir bar, 3.9a (0.95 g, 1.06 mmol, 1.2 equiv.) was

dissolved in 20mL of dry toluene at 60°C. A solution of PNPC-activated IIR (3.13)147

(2.19g, 0.882mmol of 4-nitrophenylcarbonate units, 1.0 equiv.) in 25mL of dry toluene

was added dropwise to the reaction mixture. DMAP (0.43g, 3.52mmol, 4.0 equiv.)

dissolved in 4mL dry toluene was also added to the reaction mixture, which was then

stirred overnight at 60°C. The solvent was removed in vacuo and the graft copolymer was

redissolved in DCM, washed with deionized water (3X), dried with MgSO4, concentrated

and precipitated from DCM into acetone to afford copolymer 3.14.

Yield: 2.45 g, 85% 1H NMR: δ, 5.20 (br. s, 1H), 5.10 (s, 1H), 5.05 (s, 1H), 4.87

(s, 1H), 4.24 (t, 2H, J = 5.3 Hz), 4.06 (t, 16H, J = 6.5 Hz, 3.7 (t, 2H, J = 4.7 Hz), 3.6-3.65

(m, 2H), 3.54-3.56 (m, 2H), 3.45 (q, 2H, J = 5.9 Hz), 3.38 (s, 3H), 2.53 (t, 2H, J = 6.2

Hz), 2.3 (t, 16H, J = 7.6 Hz), 1.62-1.68 (m, 32H), 1.41 (s, 88H), 1.37-1.39 (m, 16H), 1.11

(s, 264H). PCL content from 1H NMR = 15 wt%. DSC: Tg = -67°C. IR: 1165, 1230,

1366, 1390, 1470, 1736, 2955, 3445 cm-1

.

109

Graft copolymer 3.15 was prepared using the same procedure as described for polymer

3.14, expect that polymer 3.9b (1.02 g, 0.292 mmol, 0.8 equiv. relative to 4-

nitrophenylcarbonate units) was used.

Yield: 1.17 g, 70% with a conversion of 50% (based on proton integrations

corresponding to a and a’ (Figure 3.5). 1H NMR: δ, 8.27 (d, 1.5H, J = 9.1 Hz), 7.38 (d,

1.5H, J = 9.1 Hz), 5.27 (s, 0.50H), 5.20 (br. s, 0.50H), 5.12 (s, 0.50H), 5.10 (s, 0.50H),

5.05 (s, 0.50H), 4.87 (s, 0.50H), 4.13 (q, 2H, J = 7.2 Hz), 4.07 (t, 62H, J = 6.5 Hz), 3.45

(q, 2H J = 5.9 Hz), 2.53 (t, 2H, J = 5.9 Hz), 2.31 (t, 62H, J = 7.3 Hz), 1.62-1.68 (m,

124H), 1.41 (s, 88H), 1.36-1.40 (m, 62H), 1.25 (t, 3H, J = 7.3 Hz), 1.11 (s, 264H). PCL

content from 1H NMR = 32 wt%. DSC: Tg = -65°C, Tm = 44°C. IR: 1165, 1230, 1366,

1390, 1470, 1736, 2955, 3445 cm-1

.

Graft copolymer 3.16 was prepared using the same procedure as described for polymer

3.14, expect that polymer 3.9b (0.34 g, 0.2098 mmol, 1.2 equiv. relative to 4-

nitrophenylcarbonate units) was used.

Yield: 0.18 g, 75% 1H NMR: δ, 5.19 (br. s, 1H), 5.1 (s, 1H), 5.05 (br. s, 1H), 4.87

(br. s, 1H), 4.13 (q, 2H, J = 7.2 Hz), 4.07 (t, 62H, J = 6.5 Hz), 3.46 (q, 2H, J = 5.9 Hz),

2.53 (t, 2H, J = 5.9 Hz), 2.31 (t, 62H, J = 7.3 Hz), 1.62-1.68 (m, 124H), 1.41 (s, 88H),

1.36-1.40 (m, 62H), 1.25 (t, 3H, J = 7.3 Hz), 1.11 (s, 264H). PCL content from 1H NMR

= 44 wt%. DSC: Tg = -62°C, Tm = 50°C. IR: 1165, 1230, 1366, 1390, 1470, 1736, 2955,

3445 cm-1

.

Graft copolymer 3.17 was prepared using the same procedure as described for polymer

3.14, expect that polymer 3.12 (0.65 g, 0.231 mmol, 1.2 equiv. relative to 4-

nitrophenylcarbonate units) was used.

Yield: 0.51 g, 86% 1H NMR: δ, 5.14-5.25 (m, 39H), 5.11 (br. s, 1H), 5.04 (br. s,

1H), 4.86 (s, 1H), 4.23-4.32 (m, 2H), 3.57-3.59 (m, 2H), 3.45-3.50 (m, 2H), 3.36 (s, 3H),

2.61 (t, 2H, J = 2Hz), 1.54-1.59 (m, 117H), 1.41 (s, 88H), 1.11 (s, 264H). PDLLA

content from 1H NMR = 46 wt%. DSC: Tg,1 = -63°C; Tg,2 = 23°C. SEC: Mw = 485 kDa,

PDI = 3.81. IR: 1094, 1132, 1188, 1365, 1388, 1468, 1757, 2896, 2952 cm-1

.

110

4.3 Physical Characterization

4.3.1 Atomic Force Microscopy

Silicon wafers were cut into small pieces (~1cm2) and treated with “Piranha” solution, a

mixture of 3:1 H2SO2:H2O2 for approximately 1 hour to generate a clean, hydrophilic

oxide surface. The surface was then cleaned with deionized water, acetone and

subsequently dried overnight in a desiccator. Polymer thin films were prepared by spin-

casting 100 μL of a 3 wt% solution of the material in toluene on 1 cm2 of silicon wafer at

6000 rpm for 30 seconds. The surfaces were kept under vacuum for at least 24 hours

prior to image analysis. Surfaces were visualized by an atomic force microscope (XE-100

microscope from psia). Images were obtained by scanning surfaces in tapping mode with

rectangular-shaped silicon cantilevers with a spring constant of 48 N/m. Images were

then refined using XEI Image Processing software for SPM data by applying surface

smoothing and glitch removal.

4.3.2 Water Contact Angle

The water contact angle of the polymer film surface in air was measured by using a

sessile drop method on a KRÜSS DSA100 Drop Shape Analysis System (Hamburg,

Germany). Timing started after dosing a water droplet onto the testing surface, allowing

for an incubation period of 30 seconds for consistency. After 30 seconds, angles were

recorded via tangent analysis.

4.3.3 Scanning Electron Microscopy

SEM micrographs were obtained using a Hitachi 3400-N Variable Pressure Scanning

Electron Microscope. Images were taken at 100X and 1000X magnification utilizing

variable pressure mode to avoid sample preparation via gold sputtering techniques

(possible damage to rubber films).

4.3.4 Mechanical Testing of Graft Copolymers

A 40 mm x 5 mm x 0.3 mm (length x width x thickness) strip of polymer was cut from a

polymer film (prepared via compression-molding with Carver Model 385-OC heated

manual press) and its mechanical properties were measured on an INSTRON universal

111

testing machine 3300 series, at 25 mm/min and 25°C, in accordance with ASTM D882 –

12. For each copolymer, at least 6 samples were tested in separate analyses, and the data

reported is the calculated means.

4.3.5 Degradation Study

4.3.5.1 Preparation

Graft copolymers and control materials were compression-moulded using a Carver Model

385-OC (Carver Inc., Wabash) heated manual press into films approximately 0.35 mm in

thickness. Disks were punched out of the films with a diameter of 5mm, producing a

mass of approximately 5 mg per disk.

4.3.5.2 Procedure

The degradation was performed over a 4-month time period. 3 discs were measured at

each time point to obtain an average of 3 measurements. Samples were immersed in 1 mL

of 5M NaOH at 37°C in sealed vials and on completion of each time point, each sample

was rinsed with deionized water and placed in a vacuum oven at 37°C for 24 h. Lastly,

dried samples were weighed to determine % mass loss (as per Equation 9), analyzed by

size exclusion chromatography (SEC) and by scanning electron microscopy (SEM, 1

month intervals).

9)

Equation 9 – Where = average mass of initial 3 discs for a given time point and

= average mass of final 3 discs for a given time point.

The mass loss percentage was determined by calculating the average change in mass and

dividing this value by the initial average mass and multiplying by 100 for each sample at

each time point.

.

112

4.4 Biological Characterization

4.4.1 Cell Growth on Polymer Films

C2C12 mouse myoblast cells were maintained at 37oC and 5% CO2 in Dulbecco’s

Modified Eagle Medium (DMEM, Invitrogen) supplemented with 10% fetal bovine

serum (Invitrogen) and supplemented with 1% Glutamax (100) solution and 1%

Penstrep (100). Microscope glass cover slips (circular, 10 mm diameter) were coated

using a 30 mg/mL solution of the copolymers 2, 3, 4 and 5 in toluene by drop-casting (3

coats of 60 μL). The surfaces were placed in the wells of a 24-well (1 mL working

volume, NunclonMultidishes) plate and sterilized by submersion in 70% ethanol for

approximately 30 minutes, aspirated, then left to dry under UV light for approximately 1

hour. Next, the samples were conditioned overnight in Hank’s balanced salt solution

(HBSS). The sterilized samples were seeded with 2000 cells/surface and incubated for 3

hours to promote cell adhesion. Upon adhesion, 0.5 mL of cell culture media was added

to each well. The 24-well plate was incubated for 48 hours; samples were first washed

with pre-warmed 1X phosphate-buffered saline (PBS, pH 7.2) (Invitrogen) and

subsequently fixed with 4% paraformaldehyde in 1X PBS. Samples were washed three

times with 1X PBS and then treated with 0.5mL of acetone at -20oC for 5 minutes to

permeabilize the membrane. Next, they were washed another 3 times with 1X PBS,

stained with Alexa Fluor 568 phalloidin (20X dilution with 1X PBS, Invitrogen) for 15

minutes; 4’-6-diamidino-2-phenylindole dyhydrocholoride (DAPI, 300nM, 500X dilution

with 1X PBS, Invitrogen) was added directly to the surfaces for an additional 5 minutes.

The samples were washed a final 3 times with 1X PBS, placed face up onto glass

microscope slides, sealed with ProLong Gold Antifade Reagent (Invitrogen) and

contained with coverslips (type 1). Confocal images were obtained using a confocal laser

scanning microscope (LSM 510 Duo Vario, Carl Zeiss) using a 20x objective and

excitation wavelengths of 405 (DAPI) and 578 nm (Alexa Fluor 568 phalloidin). Cells

were counted using cellc12, MATLAB-based program. Statistical analyses (ANOVA

followed by Tukey’s honestly significant difference (HSD) post hoc method when

statistically significant) were performed using Prism software.

113

4.4.2 MTT Toxicity Assay

C2C12 mouse myoblast cells were first resuspended by use of Gibco® Trypsin/

ethylenediaminetetraacetic acid (EDTA) (solution of 0.025% trypsin and 0.01% EDTA in

phosphate buffered saline); a cell count was performed with a haemocytometer. Based on

this count, cells were seeded in a Nunclon® 96-well U bottom transparent polystrol plate

to obtain 10 000 cells/well in a volume of 100μL Dulbecco’s Modified Eagle Medium

(Invitrogen) supplemented with 10% fetal bovine serum (Invitrogen), 1% Glutamax

(100) solution and 1% Penstrep (100) to columns labelled C1-C8, SDS, 1-4 and VC

(vehicle control). 100μL of DMEM was introduced to the column labeled JM (just

media). The 96-well plate was then placed in a 5% CO2 incubator at 37°C for 24 hours.

Testing samples were melt-pressed (0.015") to obtain uniformity (and to provide more

surface area for potentially toxic leachates) and cut into squares of 1cm x 1cm. Samples

were then sterilized in 70% ethanol for approximately 30 minutes and subsequently dried

for 2 hours under UV light. Afterwards, they were placed in Petri dishes; 2 mL of DMEM

was added and the resulting Petri dishes were placed in a CO2 incubator at 37°C for 24

hours. After 24 hours, anything that was water soluble would be leached out of the testing

samples; DMEM that samples were subjected to was then acquired via syringe and

passed through a filter (to prevent particles >200 nm). Next, growth media from the

Nunclon® 96-well plate was aspirated. 1200 μL of the filtered DMEM testing sample

was obtained; 100μL was placed in each C8 well (as depicted in Table 4.1) to obtain a

100% leachate testing column. 600 μL of fresh DMEM was then added to the remaining

600 μL of filtered DMEM testing sample to create a 50% leachate; 100 μL was then

added to each of the wells labeled C7. This process was repeated until C1 was reached

with a 0.78% leachate dilution was attained. As a positive control to confirm cell death,

sodium lauryl sulfate (SDS) was added to the columns labeled SDS1, SDS2, SDS3 and

SDS4 with SDS concentrations in DMEM of 0.2 mg/mL, 0.15 mg/mL, 0.10 mg/mL and

0.05mg/mL, respectively. Lastly, the columns labeled VC and media (M) had fresh

DMEM introduced. The Nunclon® 96-well plate was then returned to a 5% CO2

incubator at 37°C for 24 hours. MTT (3-(4,5-Dimethylthiazol-2-yl)-2,5-

Diphenyltetrazolium Bromide) reagent was prepared at 5mg/mL in deionized water; 1

mL of this solution was added to 10 mL of DMEM. Media was aspirated from the 96-

114

well plate and replaced with 110 μL of the prepared MTT solution. The 96-well plate was

returned to the incubator for an additional 4 hours. Lastly, the MTT solution was

aspirated and replaced with 50 μL of spectroscopy grade dimethyl sulfoxide (DMSO) to

solubilize reacted product crystals. The plate was subjected to M1000-Pro (plate reader)

using Tecan i-control software at a wavelength of 540 nm. Samples were shook linearly

for 1 s with amplitude of 2 mm and a frequency of 654 rpm prior to 25 flashes.

Table 4.1 – Modified protocol adapted from ISO 10993 Technical Committee.

(2007). International Organization for Standardization (ISO)

SDS3 b b b b b b b b b b SDS1

SDS3 M C1 C2 C3 C4 C5 C6 C7 C8 VC SDS1

SDS3 M C1 C2 C3 C4 C5 C6 C7 C8 VC SDS1

SDS3 M C1 C2 C3 C4 C5 C6 C7 C8 VC SDS1

SDS4 M C1 C2 C3 C4 C5 C6 C7 C8 VC SDS2

SDS4 M C1 C2 C3 C4 C5 C6 C7 C8 VC SDS2

SDS4 M C1 C2 C3 C4 C5 C6 C7 C8 VC SDS2

SDS4 b b b b b b b b b b SDS2

115

Chapter 5

5 Conclusions and Future Work

IIR is currently used for a variety of applications ranging from foodstuffs (chewing gum),

bladders of sporting goods and most notably, as the innerliner for automotive tires.

However, in recent years there has been significant interest in expanding its application in

biomedical areas. Because of IIRs high elasticity and low strength, it has the inherent

ability to mimic various soft tissues in the body. However, in order to use it in such

applications as vascular prosthetics or orthopaedic implants, its properties have to be

tuned in order to ensure the maintenance of its structural integrity. Therefore this thesis

focused on synthesizing IIR-aliphatic polyester graft copolymers in order to tune IIR’s

chemical, physical and biological properties.

First, copolymer synthesis involving a “grafting-from” approach was investigated.

Performing the ROP of PCL from hydroxyl moieties on the IIR backbone proved to be

problematic and irreproducible. Therefore, the “grafting-to” method involving

functionalization of both the IIR backbone as well as PCL and PDLLA homopolymers

resulted in near 100% functionalization of IP units under mild conditions. Upon

successful synthesis of copolymers, they were chemically (1H NMR, SEC, DSC and

FTIR), physically (AFM, WCA, strain-controlled fatigue testing and accelerated

degradability) and biologically (cell growth on polymer films and cytoxicity testing)

characterized.

By studying the aforementioned characterization data, some important property

changes were imparted by grafting PCL or PDLLA to the IIR backbone. First, changes in

thermal properties upon copolymer preparation suggested that physical properties may be

altered. In fact, by increasing PCL/PDLLA wt%, it was found to induce phase separation,

which was confirmed via AFM analysis. By increasing the percentage of PCL from 32

wt% (copolymer 3.15) to 44 wt% (copolymer 3.16), properties were significantly altered,

as uniaxial tensile testing results showed a dramatic increase in copolymer strength. In

addition, IIR-PDLLA graft copolymer 3.17 (30 wt% PDLLA) also exhibited a

comparative (IIR) increase in UTS. The grafting of these aliphatic polyesters appeared to

116

provide physical crosslinks, thereby increasing their strength while still maintaining

compliance and structural integrity. These properties are especially important for

applications involving implantable biomaterials. In addition, grafting of polyesters may

supply ‘green strength,’ which is a measure of mechanical performance prior to

vulcanization, which is associated with harsh chemical conditions.

Copolymers consisting of lower PCL contents (copolymers 3.14 and 3.15) were

interesting in terms of increased degradability that resulted from decreased levels of

crystallinity. Long-term implantable devices that do not degrade at all elicit inflammation

due to incompatibilities with the biological host. Perhaps such limitations could be

remedied for extended-period applications through prolonged degradation and subsequent

excretion of PCL from the body. In addition, imparting degradability to oxidatively,

hydrolytically and enzymatically stable IIR may provide an avenue toward synthesis of

environmentally friendly applications.

Overall, this thesis provides an important contribution, establishing the

groundwork for understanding the properties of IIR-PCL/PDLLA graft copolymers.

Thorough analysis of chemical structure and its relation to property modification has led

to the discovery of materials with potential for usage outside the breadth of typical IIR-

based applications. Future in vivo and rheological studies will be used to elucidate their

biocompatibility and mechanical properties, respectively. These aspects require critical

analysis before copolymers can be successfully employed in biomedical applications.

117

References

(1) Langer, R. S.; Peppas, N. A. Biomaterials 1981, 2, 201.

(2) Peppas, N. A.; Langer, R. Science 1994, 263, 1715.

(3) Langer, R. Nature 1998, 392, 5.

(4) Langer, R. Chemical Engineering Science 1995, 50, 4109.

(5) Langer, R.; Tirrell, D. A. Nature 2004, 428, 487.

(6) Langer, R.; Cima, L. G.; Tamada, J. A.; Wintermantel, E. Biomaterials

1990, 11, 738.

(7) Legge, N. R.; Holden, G.; Schroeder, H. E.; Editors Thermoplastic

Elastomers. A Comprehensive Review; Hanser Publ., 1987.

(8) Sionkowska, A. Progress in Polymer Science 2011, 36, 1254.

(9) Nair, L. S.; Laurencin, C. T. Progress in Polymer Science 2007, 32, 762.

(10) Altman, G. H.; Diaz, F.; Jakuba, C.; Calabro, T.; Horan, R. L.; Chen, J.;

Lu, H.; Richmond, J.; Kaplan, D. L. Biomaterials 2003, 24, 401.

(11) Huynh, T.; Abraham, G.; Murray, J.; Brockbank, K.; Hagen, P. O.;

Sullivan, S. Nature Biotechnology 1999, 17, 1083.

(12) Narotam, P. K.; José, S.; Nathoo, N.; Taylon, C.; Vora, Y. Spine. 2004,

29, 2861.

(13) Bloomfield, S. E.; Miyata, T.; Dunn, M. W.; Bueser, N.; Stenzel, K. H.;

Rubin, A. L. Archives of Ophthalmology 1978, 96, 885.

(14) Rubin, A. L.; Stenzel, K. H.; Miyata, T.; White, M. J.; Dunn, M. Journal

of Clinical Pharmacology 1973, 13, 309.

(15) Pleyer, U.; Elkins, B.; Ruckert, D.; Lutz, S.; Grammer, J.; Chou, J.;

Schmidt, K. H.; Mondino, B. J. Current Eye Research 1994, 13, 177.

(16) Ponticiello, M. S.; Schinagl, R. M.; Kadiyala, S.; Barry, F. P. Journal of

Biomedical Materials Research 2000, 52, 246.

(17) Shantha, K. L.; Rao, K. P. Journal of Bioactive and Compatible Polymers

1993, 8, 142.

(18) Jeyanthi, R.; Panduranga Rao, K. Biomaterials 1990, 11, 238.

(19) Marty, J. J.; Oppenheim, R. C.; Speiser, P. Pharmaceutica Acta Helvetiae

1978, 53, 17.

(20) Müller, R. H.; Mäder, K.; Gohla, S. European Journal of Pharmaceutics

and Biopharmaceutics 2000, 50, 161.

(21) Bell, E.; Sher, S.; Hull, B.; Merrill, C.; Rosen, S.; Chamson, A.;

Asselineau, D.; Dubertret, L.; Coulomb, B.; Lapiere, C.; Nusgens, B.; Neveux, Y.

Journal of Investigative Dermatology 1983, 81, 2S.

(22) Boyce, S. T.; Goretsky, M. J.; Greenhalgh, D. G.; Kagan, R. J.; Rieman,

M. T.; Warden, G. D. Annals of Surgery 1995, 222, 743.

(23) Lee, C. H.; Singla, A.; Lee, Y. International Journal of Pharmaceutics

2001, 221, 1.

(24) Weinberg, C. B.; Bell, E. Science 1986, 231, 397.

(25) Freyman, T. M.; Yannas, I. V.; Gibson, L. J. Progress in Materials

Science 2001, 46, 273.

(26) O'Brien, F. J.; Harley, B. A.; Yannas, I. V.; Gibson, L. J. Biomaterials

2005, 26, 433.

118

(27) Seol, Y. J.; Lee, J. Y.; Park, Y. J.; Lee, Y. M.; Ku, Y.; Rhyu, I. C.; Lee, S.

J.; Han, S. B.; Chung, C. P. Biotechnology Letters 2004, 26, 1037.

(28) Takahashi, Y.; Yamamoto, M.; Tabata, Y. Biomaterials 2005, 26, 4856.

(29) Terbojevich, M.; Cosani, A.; Conio, G.; Marsano, E.; Bianchi, E.

Carbohydrate Research 1991, 209, 251.

(30) Borchard, G. Advanced Drug Delivery Reviews 2001, 52, 145.

(31) Di Martino, A.; Sittinger, M.; Risbud, M. V. Biomaterials 2005, 26, 5983.

(32) Kim, T. H.; Jiang, H. L.; Jere, D.; Park, I. K.; Cho, M. H.; Nah, J. W.;

Choi, Y. J.; Akaike, T.; Cho, C. S. Progress in Polymer Science (Oxford) 2007, 32, 726.

(33) Kim, I. Y.; Seo, S. J.; Moon, H. S.; Yoo, M. K.; Park, I. Y.; Kim, B. C.;

Cho, C. S. Biotechnology Advances 2008, 26, 1.

(34) Krajewska, B. Enzyme and Microbial Technology 2004, 35, 126.

(35) Puppi, D.; Chiellini, F.; Piras, A. M.; Chiellini, E. Progress in Polymer

Science 2010, 35, 403.

(36) Campoccia, D.; Doherty, P.; Radice, M.; Brun, P.; Abatangelo, G.;

Williams, D. F. Biomaterials 1998, 19, 2101.

(37) Ramamurthi, A.; Vesely, I. Journal of Biomedical Materials Research

2002, 60, 196.

(38) Li, D.-H.; Liu, L.-M.; Tian, K.-L.; Liu, J.-C.; Fan, X.-Q.

Carbohydrohydrate Polymers 2007, 67, 40.

(39) Azab, A. K.; Orkin, B.; Doviner, V.; Nissan, A.; Klein, M.; Srebnik, M.;

Rubinstein, A. Journal of Controlled Release 2006, 111, 281.

(40) Lauto, A.; Ohebshalom, M.; Esposito, M.; Mingin, J.; Li, P. S.; Felsen, D.;

Goldstein, M.; Poppas, D. P. Biomaterials 2001, 22, 1869.

(41) Shi, C.; Zhu, Y.; Ran, X.; Wang, M.; Su, Y.; Cheng, T. Journal of Surgery

Research 2006, 133, 185.

(42) Khor, E.; Lim, L. Y. Biomaterials 2003, 24, 2339.

(43) Mao, H. Q.; Roy, K.; Troung-Le, V. L.; Janes, K. A.; Lin, K. Y.; Wang,

Y.; August, J. T.; Leong, K. W. Journal of Controlled Release 2001, 70, 399.

(44) Martinac, A.; Filipovic-Grcic, J.; Voinovich, D.; Perissutti, B.;

Franceschinis, E. International Journal of Pharmacology 2005, 291, 69.

(45) Vin, F.; Teot, L.; Meaume, S. Journal of Wound Care 2002, 11, 335.

(46) Purna, S. K.; Babu, M. Burns 2000, 26, 54.

(47) Thornton, J. F.; Rohrich, R. J. Plastic and Reconstructive Surgery 2005,

116, 677.

(48) Gruessner, U.; Clemens, M.; Pahlplatz, P. V.; Sperling, P.; Witte, J.;

Rosen, H. R. American Journal of Surgery 2001, 182, 502.

(49) Olsen, D.; Yang, C.; Bodo, M.; Chang, R.; Leigh, S.; Baez, J.;

Carmichael, D.; Perala, M.; Hamalainen, E.-R.; Jarvinen, M.; Polarek, J. Advanced Drug

Delivery Reviews 2003, 55, 1547.

(50) Weigel, P. H.; Hascall, V. C.; Tammi, M. Journal of Biolical Chemistry

1997, 272, 13997.

(51) Brekke, J. H.; Thacker, K.; CRC Press LLC: 2006, p 219.

(52) Prestwich, G. D.; Marecak, D. M.; Marecek, J. F.; Vercruysse, K. P.;

Ziebell, M. R. Journal of Controlled Release 1998, 53, 93.

119

(53) Campoccia, D.; Doherty, P.; Radice, M.; Brun, P.; Abatangelo, G.;

Williams, D. F. Biomaterials 1998, 19, 2101.

(54) Cui, W.; Zhu, X.; Yang, Y.; Li, X.; Jin, Y. Materials Science and

Engineering C 2009, 29, 1869.

(55) Ji, C.; Annabi, N.; Hosseinkhani, M.; Sivaloganathan, S.; Dehghani, F.

Acta Biomaterialia 2012, 8, 570.

(56) Hong, Y.; Ye, S.-H.; Pelinescu, A. L.; Wagner, W. R. Biomacromolecules

2012, 13, 3686.

(57) Vert, M. Biomacromolecules 2005, 6, 538.

(58) Drobny, J. G. Handbook of thermoplastic elastomers; PDL(Plastics

Design Library)/William Andrew Pub: Norwich, NY, 2007.

(59) Holden, G.; John Wiley & Sons, Inc.: 2007; Vol. 24, p 695.

(60) Schollenberger, C. S.; Scott, H.; Moore, G. R. Rubber World 1958, 137,

549.

(61) Storey, R. F.; Wiggins, J. S.; Puckett, A. D. Journal of Polymer Science,

Part A: Polymer Chemistry 1994, 32, 2345.

(62) Zhang, J.-Y.; Doll, B. A.; Beckman, E. J.; Hollinger, J. O. Tissue

Engineering and Regenerative Medecine 2003, 9, 1143.

(63) Guelcher, S. A. Tissue Engineering, Part B 2008, 14, 3.

(64) Gogolewski, S.; Gorna, K. Journal of Biomedical Materials Research,

Part A 2006, 80A, 94.

(65) Laschke, M. W.; Strohe, A.; Menger, M. D.; Alini, M.; Eglin, D. Acta

Biomaterialia 2010, 6, 2020.

(66) Kavlock, K. D.; Pechar, T. W.; Hollinger, J. O.; Guelcher, S. A.;

Goldstein, A. S. Acta Biomaterialia 2007, 3, 475.

(67) Laschke, M. W.; Strohe, A.; Scheuer, C.; Eglin, D.; Verrier, S.; Alini, M.;

Pohlemann, T.; Menger, M. D. Acta Biomaterialia 2009, 5, 1991.

(68) Heller, J. Biomaterials 1980, 1, 51.

(69) Heller, J. Advanced Polymer Science 1993, 107, 41.

(70) Heller, J.; Barr, J.; Ng, S. Y.; Abdellauoi, K. S.; Gurny, R. Advanced Drug

Delivery Reviews 2002, 54, 1015.

(71) Heller, J.; Barr, J. Biomacromolecules 2004, 5, 1625.

(72) Schwach-Abdellaoui, K.; Vivien-Castioni, N.; Gurny, R. European

Journal of Pharmacology Biopharmacology 2000, 50, 83.

(73) Okada, M. Progress in Polymer Science 2002, 27, 87.

(74) Vert, M.; Christel, P.; Chabot, F.; Leray, J.; CRC: 1984, p 119.

(75) Webb, A. R.; Yang, J.; Ameer, G. A. Expert Opinion on Biological Theory

2004, 4, 801.

(76) Hayashi, T. Progress in Polymer Science 1994, 19, 663.

(77) Pistner, H.; Bendix, D. R.; Muhling, J.; Reuther, J. F. Biomaterials 1993,

14, 291.

(78) Ali, S. A. M.; Doherty, P. J.; Williams, D. F. Journal of Biomedical

Materials Research 1993, 27, 1409.

(79) Engelberg, I.; Kohn, J. Biomaterials 1991, 12, 292.

(80) Storey, R. F.; Sherman, J. W. Macromolecules 2002, 35, 1504.

120

(81) Rathi, S. R.; Coughlin, E. B.; Hsu, S. L.; Golub, C. S.; Ling, G. H.;

Tzivanis, M. J. Polymer 2012, 53, 3008.

(82) Woodruff, M. A.; Hutmacher, D. W. Progress in Polymer Science 2010,

35, 1217.

(83) Shao, J.; Sun, J. R.; Bian, X. C.; Cui, Y.; Li, G.; Chen, X. S. Journal of

Physical Chemistry B 2012, 116, 9983.

(84) Woodruff, M. A.; Hutmacher, D. W. Progress in Polymer Science

(Oxford) 2010, 35, 1217.

(85) Maurus, P. B.; Kaeding, C. C. Operative Techniques in Sports Medicine

2004, 12, 158.

(86) Fischer, E. W.; Sterzel, H. J.; Wegner, G. Kolloid-Zeitschrift & Zeitschrift

für Polymere 1973, 251, 980.

(87) Chu, C. C. J. Appl. Polym. Sci. 1981, 26, 1727.

(88) Fukushima, K.; Luis Feijoo, J.; Yang, M.-C. European Polymer Journal

2013, 49, 706.

(89) Pitt, C. G.; Gratzl, M. M.; Kimmel, G. L.; Surles, J.; Schindler, A.

Biomaterials 1981, 2, 215.

(90) Christopher, X. F. L.; Monica, M. S.; Swee-Hin, T.; Dietmar, W. H.

Biomedical Materials 2008, 3, 034108.

(91) Lee, K. H.; Kim, H. Y.; Khil, M. S.; Ra, Y. M.; Lee, D. R. Polymer 2003,

44, 1287.

(92) Luciani, A.; Coccoli, V.; Orsi, S.; Ambrosio, L.; Netti, P. A. Biomaterials

2008, 29, 4800.

(93) Thomas, V.; Jose, M. V.; Chowdhury, S.; Sullivan, J. F.; Dean, D. R.;

Vohra, Y. K. Journal of Biomaterial Science, Polymer Edition 2006, 17, 969.

(94) Huang, Q.; Goh, J. C. H.; Hutmacher, D. W.; Lee, E. H. Tissue Eng. 2002,

8, 469.

(95) Li, W. J.; Danielson, K. G.; Alexander, P. G.; Tuan, R. S. Journal of

Biomedical Materials Research - Part A 2003, 67, 1105.

(96) Kazimoǧlu, C.; Bölükbaşi, S.; Kanatli, U.; Şenköylü, A.; Altun, N. Ş.;

Babaç, C.; Yavuz, H.; Pişkin, C. International Journal of Artificial Organs 2003, 26, 804.

(97) Ajili, S. H.; Ebrahimi, N. G.; Soleimani, M. Acta Biomaterialia 2009, 5,

1519.

(98) Jeong, S. I.; Kim, B. S.; Kang, S. W.; Kwon, J. H.; Lee, Y. M.; Kim, S.

H.; Kim, Y. H. Biomaterials 2004, 25, 5939.

(99) Ananta, M.; Aulin, C. E.; Hilborn, J.; Aibibu, D.; Houis, S.; Brown, R. A.;

Mudera, V. Tissue Engineering - Part A 2009, 15, 1667.

(100) Den Dunnen, W. F. A.; Van Der Lei, B.; Schakenraad, J. M.; Stokroos, I.;

Blaauw, E.; Bartels, H.; Pennings M, A. J.; Robinson, P. H. Microsurgery 1997, 17, 348.

(101) Hakkarainen, M. Advanced Polymer Science 2002, 157, 113.

(102) Ray, S. S.; Okamoto, M. Macromolecular Materials and Engineering

2003, 288, 936.

(103) Newman, D.; Laredo, E.; Bello, A.; Grillo, A.; Feijoo, J. L.; Muller, A. J.

Macromolecules 2009, 42, 5219.

(104) Hakkarainen, M.; Albertsson, A. C.; Karlsson, S. Polymer Degradation

and Stability 1996, 52, 283.

121

(105) Reed, A. M.; Gilding, D. K. Polymer 1981, 22, 494.

(106) Hoglund, A.; Hakkarainen, M.; Albertsson, A. C. Journal of

Macromolecular Science Part a-Pure and Applied Chemistry 2007, 44, 1041.

(107) Puskas, J. E.; Chen, Y. H. Biomacromolecules 2004, 5, 1141.

(108) Puskas, J. E.; Chen, Y. H.; Dahman, Y.; Padavan, D. Journal of Polymer

Science Part A-Polymer Chemistry 2004, 42, 3091.

(109) Morton, M.; Editor Rubber Technology. 3rd Ed; Van Nostrand Reinhold,

1987.

(110) Kroschwitz, J. I. Encyclopedia of Polymer Science and Engineering 1985,

Vol. 8; Wiley, 1987.

(111) Kennedy, J. P. Rubber Chemistry and Technology 1996, 69, G63.

(112) Kennedy, J. P. Cationic Polymerization of Olefins: A Critical Inventory;

Wiley-Interscience, 1975.

(113) Young, R. J. Introduction to Polymers; Methuen, Inc., 1981.

(114) Rehner Jr, J.; Holowchak, J. Industrial and Engineering Chemistry 1944,

16, 98.

(115) Matyjaszewski, K.; Sigwalt, P. Polymer International 1994, 35, 1.

(116) Odian, G. Principles of Polymerization, 4th Edition; Wiley, 2004.

(117) Cadieux, P.; Watterson, J. D.; Denstedt, J.; Harbottle, R. R.; Puskas, J.;

Howard, J.; Gan, B. S.; Reid, G. Colloids and Surfaces B: Biointerfaces 2003, 28, 95.

(118) Pinchuk, L.; Wilson, G. J.; Barry, J. J.; Schoephoerster, R. T.; Parel, J.-M.;

Kennedy, J. P. Biomaterials 2008, 29, 448.

(119) Puskas, J. E.; Dos Santos, L. M.; Orlowski, E. Rubber Chemistry and

Technology 2010, 83, 235.

(120) Van Amerongen, G. J. Journal of Applied Physics 1946, 17, 972.

(121) Roland, C. M.; Ngai, K. L.; Plazek, D. J. Computational and Theoretical

Polymer Science 1997, 7, 133.

(122) Puskas, J. E.; Gergely, A. L.; Kaszas, G. Journal of Polymer Science Part

a-Polymer Chemistry 2013, 51, 29.

(123) Paulo, C.; Puskas, J. E. Macromolecules 2001, 34, 734.

(124) Crawford, R. A.; Morrissey, R. T.; B. F. Goodrich Co.: 1953.

(125) Morrissey, R. T. Journal of Industrial Engineering Chemistry 1955, 47,

1562.

(126) Morrissey, R. T.; Weiss, H. J.; B. F. Goodrich Co.: 1958.

(127) Baldwin, F. P.; Thomas, R. M.; Esso Research and Engineering Co.: 1960.

(128) Baldwin, F. P.; Thomas, R. M.; Esso Research and Engineering Co.: 1960.

(129) Walker, J.; Jones, R. H.; Feniak, G. Rubber Age (NY) 1976, 108, 45.

(130) Edwards, D. C. Elastomerics 1990, 122, 19.

(131) Kuntz, I.; Zapp, R. I.; Pancirov, R. J. Rubber Chemistry and Technology

1984, 57, 813.

(132) McLean, J. K.; Guillen-Castellanos, S. A.; Parent, J. S.; Whitney, R. A.;

Resendes, R. European Polymer Journal 2007, 43, 4619.

(133) Xiao, S.; Parent, J. S.; Whitney, R. A.; Knight, L. K. Journal of Polymer

Science, Part A: Polymer Chemistry 2010, 48, 4691.

(134) Malmberg, S. M.; Parent, J. S.; Pratt, D. A.; Whitney, R. A.

Macromolecules 2010, 43, 8456.

122

(135) Whitney, R. A.; Penciu, A.; Parent, J. S.; Resendes, R.; Hopkins, W.

Macromolecules 2005, 38, 4625.

(136) Parent, J. S.; White, G. D. F.; Thom, D. J.; Whitney, R. A.; Hopkins, W.

Journal of Polymer Science, Part A: Polymer Chemistry 2003, 41, 1915.

(137) Guillén-Castellanos, S. A.; Parent, J. S.; Whitney, R. A. Macromolecules

2006, 39, 2514.

(138) Puskas, J. E.; Antony, P.; El Fray, M.; Altstädt, V. European Polymer

Journal 2003, 39, 2041.

(139) Kennedy, J. P.; Richard, G. C. Macromolecules 1993, 26, 567.

(140) Keszler, B.; Fenyvesi, G. Y.; Kennedy, J. P. Journal of Polymer Science,

Part A: Polymer Chemistry 2000, 38, 706.

(141) Feldthusen, J.; Iván, B.; Müller, A. H. E. Macromolecules 1998, 31, 578.

(142) Ojha, U.; Kulkarni, P.; Cozzens, D.; Faust, R. Journal of Polymer Science

Part A-Polymer Chemistry 2010, 48, 3767.

(143) Gong, C.; Fréchet, J. M. J. Journal of Polymer Science, Part A: Polymer

Chemistry 2000, 38, 2970.

(144) Groenewolt, M.; Brezesinski, T.; Schlaad, H.; Antonietti, M.; Groh, P. W.;

Iván, B. Advanced Materials 2005, 17, 1158.

(145) Zhang, Y.; Zhang, Y.; Chen, X. Z.; Zhang, Y. Reactive and Functional

Polymers 2001, 47, 93.

(146) Zhang, Y.; Chen, X.; Zhang, Y.; Zhang, Y. Macromolecular Materials

and Engineering 2001, 286, 443.

(147) Bonduelle, C. V.; Gillies, E. R. Macromolecules 2010, 43, 9230.

(148) Bonduelle, C. V.; Karamdoust, S.; Gillies, E. R. Macromolecules 2011,

44, 6405.

(149) Suksawad, P.; Yamamoto, Y.; Kawahara, S. European Polymer Journal

2011, 47, 330.

(150) Pukkate, N.; Yamamoto, Y.; Kawahara, S. Colloid and Polymer Science

2008, 286, 411.

(151) Bhakuni, R. S.; Mowdood, S. K.; Waddell, W. H.; Rai, I. S.; Knight, D.

L.; Wiley: 1989; Vol. 16, p 834.

(152) Dunn, J. R. Elastomerics 1991, 123, 15.

(153) Trexler, H. E.; Lee, M. C. H. Journal of Applied Polymer Science 1986,

32, 3899.

(154) Wong, W. K. Rubber World 2009, 240, 20.

(155) Pinchuk, L.; Wilson, G. J.; Barry, J. J.; Schoephoerster, R. T.; Parel, J. M.;

Kennedy, J. P. Biomaterials 2008, 29, 448.

(156) Puskas, J. E.; Chen, Y. Biomacromolecules 2004, 5, 1141.

(157) Acosta, A. C.; Espana, E. M.; Yamamoto, H.; Davis, S.; Pinchuk, L.;

Weber, B. A.; Orozco, M.; Dubovy, S.; Fantes, F.; Parel, J. M. Archives of

Ophthalmology 2006, 124, 1742.

(158) Gallocher, S. L.; Aguirre, A. F.; Kasyanov, V.; Pinchuk, L.;

Schoephoerster, R. T. Journal of Biomedical Materials Research - Part B Applied

Biomaterials 2006, 79, 325.

(159) Kennedy, J. P.; Askew, M. J.; Richard, G. C. Journal of biomaterials

science. Polymer edition 1993, 4, 445.

123

(160) Kennedy, J. P. Macromolecular Symposia 2001, 175, 127.

(161) Kennedy, J. P.; Midha, S.; Gadkari, A. Journal of Macromolecular

Science-Chemistry 1991, A28, 209.

(162) Isayeva, I. S.; Kasibhatla, B. T.; Rosenthal, K. S.; Kennedy, J. P.

Biomaterials 2003, 24, 3483.

(163) Andrade, J. D.; Hlady, V.; Jeon, S. I. Advances in Chemistry Series 1996,

248, 57.

(164) Puskas, J. E.; Kwon, Y.; Antony, P.; Bhowmick, A. K. Journal of Polymer

Science Part a-Polymer Chemistry 2005, 43, 1811.

(165) Aguilar, M.; Yamamoto, H.; Espana, E.; Acosta, A. C.; Orozco, M.; Aly,

M.; Arrieta, E.; Hernandez, E.; Martin, J.; Dubovy, S.; Smiddy, W.; Pinchuk, L.; Parel,

J.-M. Proceedings of SPIE - The International Society for Optical Engineering 2006,

6138, 61381S/1.

(166) Perez-Santonja, J. J.; Alio, J. L.; Jimenez-Alfaro, I.; Zato, M. A. Journal

of Cataract and Refractive Surgery 2000, 26, 1288.

(167) Gallocher, S. L.; Aguirre, A. F.; Kasyanov, V.; Pinchuk, L.;

Schoephoerster, R. T. Journal of Biomedical Materials Research Part B-Applied

Biomaterials 2006, 79B, 325.

(168) Yin, W.; Gallocher, S.; Pinchuk, L.; Schoephoerster, R. T.; Jesty, J.;

Bluestein, D. Artificial Organs 2005, 29, 826.

(169) Levy, Y.; Mandler, D.; Weinberger, J.; Domb, A. J. Journal of Biomedical

Materials Research - Part B Applied Biomaterials 2009, 91, 441.

(170) Ormiston, J. A.; Webster, M. W.; Ruygrok, P. N.; Stewart, J. T.; Currie,

E.; Panther, M. J. American Journal of Cardiology 2002, 90, 76H.

(171) Otsuka, Y.; Chronos, N. A.; Apkarian, R. P.; Robinson, K. A. The Journal

of invasive cardiology 2007, 19, 71.

(172) Duraiswamy, N.; Choksi, T. D.; Pinchuk, L.; Schoephoerster, R. T.

Journal of Biomaterials Applications 2009, 23, 367.

(173) Janssen, D.; De Palma, R.; Verlaak, S.; Heremans, P.; Dehaen, W. Thin

Solid Films 2006, 515, 1433.

(174) Frommer, J. Angewandte Chemie-International Edition 1992, 104, 1325.

(175) Magonov, S. N.; Whangbo, M. H.; Editors Surface Analysis with STM and

AFM: Experimental and Theoretical Aspects of Image Analysis; VCH, 1995.

(176) Magonov, S. N.; Elings, V.; Whangbo, M. H. Surface Science 1997, 375,

L385.

(177) Van, N. S. J. T.; Van, d. W. K. O.; De, G. B. G.; Van, H. N. F.; Greve, J.

Ultramicroscopy 1997, 69, 117.

(178) Magonov, S. N.; Cleveland, J.; Elings, V.; Denley, D.; Whangbo, M. H.

Surface Science 1997, 389, 201.

(179) Haward, R.; Young, R. J.; Editors The Physics of Glassy Polymers,

Second Edition; Chapman & Hall, 1997.

(180) Bassett, D. C. Principles of Polymer Morphology; Cambridge Univ. Press,

1981.

(181) Schultz, J. Polymer Materials Science; Prentice-Hall, 1973.

(182) Bartczak, Z.; Galeski, A. Macromolecular Symposia 2010, 294, 67.

(183) Boyer, R. F. Rubber Chemistry Technology 1963, 36, 1303.

124

(184) Legge, N. R.; Holden, G.; Schroeder, H. E. Thermoplastic elastomers: a

comprehensive review; Hanser Publishers, 1987.

(185) Semon, W. L.; B. F. Goodrich Co. . 1933.

(186) Perego, G.; Cesari, M.; Vitali, R. Journal of Applied Polymer Science

1984, 29, 1157.

(187) Birley, A. W. British Polymer Journal 1989, 21, 181.

(188) Callister, W. D., Jr. Fundamentals of Materials Science and Engineering:

An Interactive E-Text, 2nd Edition; Wiley, 2004.

(189) Ashby, M. F.; Ferreira, P. J.; Schodek, D. L. In Nanomaterials,

Nanotechnologies and Design; Butterworth-Heinemann: Boston, 2009, p 87.

(190) Albert, D.; Hoffmann, A. Medical Device and Diagnostic Industry 2008,

30.

(191) Pizzoferrato, A.; Ciapetti, G.; Stea, S.; Cenni, E.; Arciola, C. R.; Granchi,

D.; Savarino, L. Clinical Materials 1994, 15, 173.

(192) Ratner, B. D.; Hoffman, A. S.; Schoen, F. J.; Lemons, J. E.; Editors

Biomaterials Science, An Introduction to Materials in Medicine, 2nd Edition; Elsevier,

2004.

(193) Chen, H.; Yuan, L.; Song, W.; Wu, Z.; Li, D. Progress in Polymer Science

2008, 33, 1059.

(194) Ryoo, S.-R.; Kim, Y.-K.; Kim, M.-H.; Min, D.-H. ACS Nano 2010, 4,

6587.

(195) Shin, Y. M.; Kim, K.-S.; Lim, Y. M.; Nho, Y. C.; Shin, H.

Biomacromolecules 2008, 9, 1772.

(196) Huang, J.; Grater, S. V.; Corbellini, F.; Rinck, S.; Bock, E.; Kemkemer,

R.; Kessler, H.; Ding, J.; Spatz, J. P. Nanoscience and Nanotechnolgy Letters 2009, 9,

1111.

(197) Discher, D. E.; Janmey, P.; Wang, Y.-l. Science 2005, 310, 1139.

(198) Brunetti, V.; Maiorano, G.; Rizzello, L.; Sorce, B.; Sabella, S.; Cingolani,

R.; Pompa, P. P. Proceedings of the National Academy of Sciences of the U. S. A. 2010,

107, 6264.

(199) Cheng, Z.; Teoh, S.-H. Biomaterials 2004, 25, 1991.

(200) Dugan, J. M.; Gough, J. E.; Eichhorn, S. J. Biomacromolecules 2010, 11,

2498.

(201) Huang, N. F.; Patel, S.; Thakar, R. G.; Wu, J.; Hsiao, B. S.; Chu, B.; Lee,

R. J.; Li, S. Nanoscience and Nanotechnology Letters 2006, 6, 537.

(202) Lampin, M.; Warocquier-Clérout, R.; Legris, C.; Degrange, M.; Sigot-

Luizard, M. F. Journal of Biomedical Materials Research 1997, 36, 99.

(203) Brash, J. L.; Horbett, T. A. ACS Symposium Series 1995, 602, 1.

(204) de Vos, W. M.; de Keizer, A.; Kleijn, J. M.; Cohen Stuart, M. A.

Langmuir 2009, 25, 4490.

(205) Norde, W.; Gage, D. Langmuir 2004, 20, 4162.

(206) Aulich, D.; Hoy, O.; Luzinov, I.; Brücher, M.; Hergenröder, R.; Bittrich,

E.; Eichhorn, K. J.; Uhlmann, P.; Stamm, M.; Esser, N.; Hinrichs, K. Langmuir 2010, 26,

12926.

(207) Uhlmann, P.; Houbenov, N.; Brenner, N.; Grundke, K.; Burkert, S.;

Stamm, M. Langmuir 2007, 23, 57.

125

(208) Jeon, S. I.; Lee, J. H.; Andrade, J. D.; De Gennes, P. G. Journal of Colloid

And Interface Science 1991, 142, 149.

(209) Taunton, H. J.; Toprakcioglu, C.; Fetters, L. J.; Klein, J.; 1 ed. 1989; Vol.

30, p 368.

(210) Unsworth, L. D.; Sheardown, H.; Brash, J. L. Biomaterials 2005, 26,

5927.

(211) Harder, P.; Grunze, M.; Dahint, R.; Whitesides, G. M.; Laibinis, P. E.

Journal of Physical Chemistry B 1998, 102, 426.

(212) Yamashita, S.; Kodama, K.; Ikeda, Y.; Kohjiya, S. Journal of Polymer

Science Part A: Polymer Chemistry 1993, 31, 2437.

(213) Bonduelle, C. V.; Lau, W. M.; Gillies, E. R. ACS Applied Materials &

Interfaces 2011, 3, 1740.

(214) Karamdoust, S.; Bonduelle, C. V.; Amos, R. C.; Turowec, B. A.; Guo, S.;

Ferrari, L.; Gillies, E. R. Journal of Polymer Science Part A: Polymer Chemistry 2013,

n/a.

(215) Vert, M.; Li, S. M.; Spenlehauer, G.; Guerin, P. Journal of Materials

Science: Materials in Medicine 1992, 3, 432.

(216) Wei, X.; Bagdi, K.; Ren, L.; Shah, P.; Seethamraju, K.; Faust, R. Polymer

2013, 54, 1647.

(217) Li, B.; Martin, A. L.; Gillies, E. R. Chemical Communications 2007, 5217.

(218) Chakrapani, V. Y.; Gnanamani, A.; Giridev, V. R.; Madhusoothanan, M.;

Sekaran, G. Journal of Applied Polymer Science 2012, 125, 3221.

(219) Fan, Y. J.; Chen, G. P.; Tanaka, J.; Tateishi, T. Biomacromolecules 2005,

6, 3051.

(220) Harrane, A.; Belaouedj, M. E. A.; Belbachir, M. Reactive and Functional

Polymers 2011, 71, 126.

(221) Zhang, P.; Huang, H.; Yan, D.; He, T. Langmuir 2012, 28, 6419.

(222) Vasilev, C.; Reiter, G.; Pispas, S.; Hadjichristidis, N. Polymer 2006, 47,

330.

(223) Nojima, S.; Nakano, H.; Takahashi, Y.; Ashida, T. Polymer 1994, 35,

3479.

(224) Hsu, J.-Y.; Hsieh, I. F.; Nandan, B.; Chiu, F.-C.; Chen, J.-H.; Jeng, U. S.;

Chen, H.-L. Macromolecules 2007, 40, 5014.

(225) Nandan, B.; Hsu, J.-Y.; Chen, H.-L. Polymer Reviews 2006, 46, 143.

(226) Gheno, S. M.; Passador, F. R.; Pessan, L. A. Journal of Applied Polymer

Science 2010, 117, 3211.

(227) Kumar, S.; Rath, T.; Mahaling, R. N.; Das, C. K.; Srivastava, R. B.;

Yadaw, S. B. Polymer Composites 2009, 30, 655.

(228) Tsou, A. H.; Duvdevani, I.; Datta, S. Journal of Applied Polymer Science

2006, 102, 4447.

(229) Corrales, T.; Larraza, I.; Catalina, F.; Portoles, T.; Ramirez-Santillan, C.;

Matesanz, M.; Abrusci, C. Biomacromolecules 2012, 13, 4247.

(230) Huang, Y.; Zhang, S.; Niu, B.; Wang, D.; Wang, Z.; Feng, S.; Xu, H.;

Kong, D.; Qiao, M. Colloids and Surfaces B: Biointerfaces 2013, 101, 361.

(231) Li, Y.; Zhang, B.; Ruan, C.; Wang, P.; Sun, J.; Pan, J.; Wang, Y. Journal

of Biomedical Materials Research - Part A 2012, 100 A, 3496.

126

(232) Martínez, L.; Nevshupa, R.; Álvarez, L.; Huttel, Y.; Méndez, J.; Román,

E.; Mozas, E.; Valdés, J. R.; Jimenez, M. A.; Gachon, Y.; Heau, C.; Faverjon, F.

Tribology International 2009, 42, 584.

(233) Bonduelle, C. V.; McEachran, M. J.; Karamdoust, S.; Gillies, E. R.

Journal of Coatings Technology Research 2013, 1.

(234) Jiang, T.; He, F.; Zhuo, R.-X. Polymer Degradation and Stability 2013,

98, 325.

(235) Chen, Q.; Liang, S.; Thouas, G. A. Progress in Polymer Science 2013, 38,

584.

(236) Balasubramanian, V.; Grusin, N. K.; Bucher, R. W.; Turitto, V. T.; Slack,

S. M. Journal of Biomedical Materials Research 1999, 44, 253.

(237) Cozzens, D.; Luk, A.; Ojha, U.; Ruths, M.; Faust, R. Langmuir 2011, 27,

14160.

(238) Harte, I.; Birkinshaw, C.; Jones, E.; Kennedy, J.; Debarra, E. Journal of

Applied Polymer Science 2013, 127, 1997.

(239) Castilla-Cortázar, I.; Más-Estellés, J.; Meseguer-Dueñas, J. M.; Escobar

Ivirico, J. L.; Marí, B.; Vidaurre, A. Polymer Degradation and Stability 2012, 97, 1241.

(240) Ji, C.; Annabi, N.; Hosseinkhani, M.; Sivaloganathan, S.; Dehghani, F.

Acta Biomaterialia 2012, 8, 570.

(241) Pok, S.; Myers, J. D.; Madihally, S. V.; Jacot, J. G. Acta Biomaterialia

2013, 9, 5630.

(242) Roy, R. V.; Das, M.; Banerjee, R.; Bhowmick, A. K. Bioresource

Technology 2006, 97, 2485.

(243) Ikeda, Y.; Kodama, K.; Kajiwara, K.; Kohjiya, S. Journal of Polymer

Science Part B-Polymer Physics 1995, 33, 387.

(244) Xu, Y.; Petrovic, Z.; Das, S.; Wilkes, G. L. Polymer 2008, 49, 4248.

(245) Kaszas, G. 1993; Vol. 68, p 325.

(246) Safranski, D. L.; Smith, K. E.; Gall, K. Polymer Reviews 2013, 53, 76.

(247) Kennedy, J. P. Macromolecular Symposia 2001, 175, 127.

(248) Van, D. J. D.; Gnatowski, M.; Burczyk, A. Journal of Applied Polymer

Science 2008, 109, 1535.

(249) Htay, A. S.; Teoh, S. H.; Hutmacher, D. W. Journal of Biomaterial

Science, Polymer Edition 2004, 15, 683.

(250) Cam, D.; Hyon, S.-H.; Ikada, Y. Biomaterials 1995, 16, 833.

(251) Li, S. M.; Chen, X. H.; Gross, R. A.; McCarthy, S. P. Journal of Materials

Science: Materials in Medicine 2000, 11, 227.

(252) Ojha, U.; Kulkarni, P.; Singh, J.; Faust, R. Journal of Polymer Science

Part A-Polymer Chemistry 2009, 47, 3490.

(253) Chen, C. M.; Yang, S. Polymer International 2012, 61, 1041.

(254) Stafford, C. M.; Harrison, C.; Beers, K. L.; Karim, A.; Amis, E. J.;

Vanlandingham, M. R.; Kim, H. C.; Volksen, W.; Miller, R. D.; Simonyi, E. E. Nature

Materials 2004, 3, 545.

(255) Kelch, S.; Choi, N. Y.; Wang, Z.; Lendlein, A. Advanced Engineering

Materials 2008, 10, 494.

(256) Lendlein, A.; Schmidt, A. M.; Langer, R. Proceedings of the National

Academy of Sciences 2001, 98, 842.

127

(257) Yakacki, C. M.; Lyons, M. B.; Rech, B.; Gall, K.; Shandas, R. Biomedical

Materials (Bristol, U. K.) 2008, 3, 015010/1.

(258) Liu, C.; Mather, P. T. Journal of Applied Medical Polymers 2002, 6, 47.

(259) Kelch, S.; Choi, N.-y.; Wang, Z.; Lendlein, A. Advanced Engineering

Materials 2008, 10, 494.

(260) Lendlein, A.; Schmidt, A. M.; Schroeter, M.; Langer, R. Journal of

Polymer Science, Part A: Polymer Chemistry 2005, 43, 1369.

(261) Smith, K. E.; Trusty, P.; Wan, B.; Gall, K. Acta Biomaterialia 2011, 7,

558.

(262) Miao, H.; Hao, J.; Liu, Y.; Liu, Y.; Deng, X. Polymer International 2008,

57, 316.

(263) Burkersroda, F. v.; Schedl, L.; Göpferich, A. Biomaterials 2002, 23, 4221.

(264) Gopferich, A. Biomaterials 1996, 17, 103.

(265) Pitt, C. G.; Chasalow, F. I.; Hibionada, Y. M.; Klimas, D. M.; Schindler,

A. Journal of Applied Polymer Science 1981, 26, 3779.

(266) Tanaka, S.; Ogura, A.; Kaneko, T.; Murata, Y.; Akashi, M.

Biomacromolecules 2004, 5, 2447.

(267) Cole, M. A.; Voelcker, N. H.; Thissen, H.; Griesser, H. J. Biomaterials

2009, 30, 1827.

(268) Chinn, J. A.; Sauter, J. A.; Phillips Jr, R. E.; Kao, W. J.; Anderson, J. M.;

Hanson, S. R.; Ashton, T. R. Journal of Biomedical Materials Research 1997, 39, 130.

128

Appendices

Appendix A – 1H NMR depicting functionalization of 3.7b: a) initial homopolymer;

b) reacted with t-BOC-protected β-alanine (3.8b); c) and deprotected with TFA

(3.9b).

129

Appendix B – 1H NMR showing PCL content of: a) copolymer 3.16 (44 wt% PCL)

and b) copolymer 3.15 (32wt% PCL).

130

Appendix C – FTIR of: a) 3.8a and 3.9a; b) 3.11 and 3.12.

131

Appendix D – FTIR of: a) copolymer 3.14; b) copolymer 3.15.

132

Appendix E – FTIR of: a) copolymer 3.16; b) copolymer 3.17.

133

Appendix F – DSC traces of: IIR-PCL copolymer after 1st precipitation; IIR-PCL

copolymer after a 2nd

precipitation into acetone, ridding of free homopolymer as

evidenced by disappearance of the second Tm at 53°C.

Appendix G – DSC trace depicting the two Tgs of copolymer 3.17.

134

Appendix H – DSC traces depicting Tg and Tm of IIR-PCL graft copolymers: a)

copolymer 3.14; b) copolymer 3.15; c) copolymer 3.16.

Appendix I – DSC traces depicting Tg and Tm of IIR-PCL/PLA blends.

135

Appendix J – Raw data coinciding with Figure 3.21 for the degradation study of

IIR-PCL copolymers and IIR control.

Appendix K – Raw data coinciding with Figure 3.22 for the degradation study of

IIR-PDLLA copolymer and IIR control.

136

Curriculum Vitae

EDUCATION

The University of Western Ontario, London, Ontario

Master of Engineering Science, Biomedical Engineering Graduate Program June 2013

Dissertation: “Butyl Rubber and Polyester Graft Copolymers for Biomedical Applications”

Supervisor: Dr. Elizabeth Gillies

The University of Western Ontario, London, Ontario

Bachelor of Engineering Science, Department of Chemical and Biochemical Engineering April 2011

Specialization: Biochemical and Environmental Engineering

Graduated with Distinction

Honours Thesis: “Utilizing Mesoporous Ordered Silicates as Catalysts for Biodiesel

Production”

EMPLOYMENT EXPERIENCE

The University of Western Ontario | London

Teaching Assistant – to Dr. Shazhad Barghi in “Introduction to Plant Design and Safety” Jan – Apr 2013

Provided student assistance for research assistance and plant development for group-dependent

processes

Evaluated student performance based on a standardized rubric

Met with students upon request to optimize plant design

Teaching Assistant – to Dr. Elizabeth Gillies in “Industrial Organic Chemistry” Sept 2012

Explained methodologies to students pertaining to successful completion of hands-on experiments

Evaluated student in-lab performance as well as comprehension of background theory based on

completion of formal reports

Available to meet with students upon request and marked course midterms

Teaching Assistant – to Dr. Shazhad Barghi in “Introductory Engineering Design and Innovation Studio” Sept 2011

Provided student assistance on the iterative development for several design-based projects

Evaluated student performance based on a standardized rubric

Met with students upon request and graded all written work

Research Assistant – to Dr. Elizabeth Gillies | Department of Chemistry May – Aug 2011

Developed synthetic laboratory skills necessary for working in a polymer research facility Worked on butyl rubber project investigating graft copolymerization to enhance material properties Optimized toxicity cell assays and imaging for biocompatibility testing

NSERC Undergraduate Student Researcher – to Dr. Sohrab Rohani | Chemical Engineering Ap/10 – Ma/11

Developed and optimized biodiesel production with a heterogeneous-based catalyst

Worked as a team member and enhanced laboratory, communication and research skills

Obtained particle characterization skills using various instruments

AWARDS

Ontario Graduate Scholarship, The University of Western Ontario September 2012

Natural Science and Engineering Research Council of Canada, Undergraduate Student Research Award May 2010

Dean’s Honour List, Faculty of Engineering Science April 2008

All Canadian/Ontario University Athletics Academic Achievement Award 2007 – 2012

Queen Elizabeth II Aiming for the Top Scholarship 2007 – 2011

The Governor General’s Academic Medal (highest overall average) June 2006

137

PUBLICATIONS AND PAPERS

Turowec, B..; Gillies, E.; Functionalized Polyisobutylene-Based Materials for Biomedical

Applications.

96th Canadian Chemistry and Exhibition, Québéc City, Québéc.

Presented in oral format. May 2013

Karamdoust, S.; Bonduelle, C.V.; Amos, R.C.; Turowec, B.A..; Guo, S.; Ferrari, L.; Gillies, E.;

Journal of Polymer Science: Part A, 2013. Synthesis and properties of butyl rubber –

poly(ethylene oxide) graft copolymers with high PEO content. May 2013

Turowec, B.; Gillies, E.R.; Polyisobutylene-based Rubber and Polyester Graft Copolymers for

Biomedical Applications.

The University of Western Ontario, Biomedical Engineering Seminar Presentation.

Presented in oral format. April 2013

Kazemian, H.; Turowec, B.; Siddiquee, M. N.; Rohani, S.; Fuel, 2013, 103, 719.

Biodiesel Production using cesium modified mesoporous ordered silica as

heterogeneous base catalyst. January 2013

Turowec, B.; Gillies, E.R.; The Expansion of Polyisobutylene into Biomedical Applications.

The University of Western Ontario, Biomedical Engineering Seminar Presentation.

Presented in oral format. April 2012

Turowec, B.; Gillies, E.R.; Developing Polyisobutylene for Biomedical Applications.

The University of Western Ontario, The London Health Research Day. Presented in

poster format. March 2012

Turowec, B.; Siddiquee, M.N.; Kazemian, H.; and Rohani, S.; Zeolitic Mesoporous Ordered Silica as

Heterogeneous Solid Catalysts for Biodiesel Production.

The University of Western Ontario, Particle Technology Research Centre

Conference, London, ON. Presented in poster format. Received 3rd prize. August 2010

MEMBERSHIPS

The University of Western Ontario | London

Programs Operation Committee 2011 – 2013

Department of Biomedical Engineering

Leader’s Circle 2011 – 2012

Mustang Varsity Athletics

VOLUNTEER ACTIVITIES

The University of Western Ontario Student Orientation Guide 2011

The Western Fair 2010 & 2011

Chemistry/Calculus Tutor 2009 – 2011

CO-CURRICULAR ACTIVITIES

Kawartha Lakes Lacrosse, Senior ‘A’ Team, Ontario 2010 – Present

Women’s Varsity Lacrosse, The University of Western Ontario 2007 – 2011

Won Ontario University Athletic Championship (2009, 2011)

Captain of Lacrosse Team

Women’s Ontario Ball Hockey Team 2004 – 2011

Won National Championship (2009)

Women’s Varsity Hockey Team, The University of Western Ontario 2006 – 2007

Lady Blue Knights Lacrosse, U19 A Team, Ontario

Won Provincial Finals 2005 – 2006